You are on page 1of 7

Carbohydrate Polymers 179 (2018) 379–385

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Quantification of food polysaccharide mixtures by 1H NMR MARK


a a a,⁎ a
Donny W.H. Merkx , Yvonne Westphal , Ewoud J.J. van Velzen , Kavish V. Thakoer ,
Niels de Rooa, John P.M. van Duynhovena,b
a
Unilever R & D, Olivier van Noortlaan 120, 3133 AT, Vlaardingen, The Netherlands
b
Wageningen University & Research, Laboratory of Biophysics, Stippeneng 4, 6708 WE, Wageningen, The Netherlands

A R T I C L E I N F O A B S T R A C T

Keywords: Polysaccharides are food ingredients that critically determine rheological properties and shelf life. A qualitative
Food polysaccharides and quantitative assessment on food-specific polysaccharide mixtures by 1H NMR is presented. The method is
Mixture analysis based on the identification of intact polysaccharides, combined with a quantitative analysis of their mono-
Non-instantaneous monosaccharide release saccharide constituents. Identification of the polysaccharides is achieved by 1H NMR line shape fitting with pure
qNMR
compound spectra. The monomeric composition was determined using the Saeman hydrolysis procedure, fol-
Saeman hydrolysis
lowed by direct monosaccharide quantification by 1H NMR. In the quantification, both the monosaccharide
degradation during hydrolysis, as well as a correction for the non-instantaneous polysaccharide dissolution were
taken into account. These factors were particularly important for the quantification of pectins. The method
showed overall good repeatability (RSDr = 4.1 ± 0.9%) and within-laboratory reproducibility
(RSDR = 6.1 ± 1.4%) for various food polysaccharides. Polysaccharide mixtures were quantitatively resolved
by a non-negative least squares estimation, using identified polysaccharides and their molar monosaccharide
stoichiometry as prior knowledge. The accuracy and precision of the presented method make it applicable to a
wide range of food polysaccharide mixtures with complex and overlapping 1H NMR spectra.

1. Introduction strong β-1,4-glycosidic bonds of for example cellulose, whilst being


relatively mild towards most released monosaccharides (Saeman et al.,
Polysaccharides are widely used in food products and critically 1954). Typically, the monosaccharide content after degradation is de-
determine rheological properties (Dickinson, 2003; termined by combining GC and colorimetric assays
Saha & Bhattacharya, 2010) and shelf life (Wang, Li, Copeland, (Ahmed & Labavitch, 1977; Englyst & Cummings, 1984). These methods
Niu, & Wang, 2015). In order to understand and manipulate these are, however, not able to quantify neutral sugars and uronic acids in a
macroscopic properties, a detailed and quantitative knowledge on the single experiment and are not able to distinguish between different
polysaccharide composition of complex food products is required. This uronic acids. Furthermore, typical quantification of neutral sugars re-
is a challenging task since food polysaccharides show a wide structural quire laborious derivatisation prior to GC analysis, which limit the
diversity, are often formulated at low levels and engage in strong in- overall sample throughput significantly.
teractions with the food matrix. Polysaccharide fractions can be readily Quantitative 1H NMR (qNMR) analysis has been introduced as a
isolated from complex food matrices, with mg-scale yields (Grün et al., high-throughput quantification method (van Duynhoven, van
2015). The quantification of polysaccharide mixtures, however, re- Velzen, & Jacobs, 2013). It does not require monosaccharide specific
mains a significant challenge. Commonly used techniques for poly- calibration and for quantification only one primary standard needs to
saccharide quantification are based on hydrolysis/methanolysis of the be used, either internal or external. qNMR can be applied directly on
polysaccharides, followed by quantification of the released mono- the hydrolysate mixture in sulphuric acid, which significantly simplifies
saccharides. sample preparation. Recently, two research groups (Carvalho de Souza,
Hydrolysis with hydrochloric acid (HCl) (Bertaud, Rietkerk, Selin, & Lankhorst, 2013; Mittal, Scott, Amidon,
Sundberg, & Holmbom, 2002), trifluoroacetic acid (TFA) (De Ruiter, Kiemle, & Stipanovic, 2009) have tailored the Saeman hydrolysis pro-
Schols, Voragen, & Rombouts, 1992) and sulphuric acid (Saeman, cedure towards 1H NMR quantification by exploring the use of deut-
Moore, Mitchell, & Millet, 1954) are the most common approaches. The erated sulphuric acid (D2SO4) and optimising the sample preparation
Saeman hydrolysis procedure uses sulphuric acid (H2SO4) to break the for 1H NMR. For quantification purposes, a degradation factor is


Corresponding author.
E-mail address: Ewoud-van.Velzen@Unilever.com (E.J.J. van Velzen).

http://dx.doi.org/10.1016/j.carbpol.2017.09.074
Received 26 June 2017; Received in revised form 20 September 2017; Accepted 22 September 2017
Available online 28 September 2017
0144-8617/ © 2017 Elsevier Ltd. All rights reserved.
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

commonly applied to correct for the degradation of the mono- 2.3. Polysaccharide hydrolysis
saccharides during the hydrolysis step (De Ruiter et al., 1992). This
factor does not take the rate of monomeric release into account and Approximately 25 mg of a polysaccharide sample was accurately
inherently assumes that all monosaccharides are instantaneously re- weighed into a 20 mL glass tube. After adding 1 mL of 72% (w/w)
leased at timepoint to. This assumption, however, may not hold for all D2SO4 solution to the sample, the glass tube was sealed and stirred at
monosaccharides. In cases where the monosaccharide release evolves room temperature for 1 h. Upon dissolution, 6.2 mL of D2O was added
over a time interval Δt > 0, the quantification will result in over- to obtain a final D2SO4 concentration of 13% (w/w) in D2O. The sample
estimated results. Moreover, the currently available NMR methods have was then incubated in a convection oven at 100 °C for 180 min. In the
been primarily optimised for analysing biomass and not to specifically specific case of the dissolution determination of polysaccharides, the
consider the compositional complexity of polysaccharides as they occur samples were incubated between 0 and 180 min, with increments of 15
in foods (Carvalho de Souza et al., 2013; Mittal et al., 2009). or 30 min. After incubation, the sample was allowed to cool down to
In the currently reported study, we optimised and validated a rapid room temperature, after which 1 mL of maleic acid solution (internal
qNMR method for the quantification of polysaccharide mixtures in standard) was added. The samples were, when required, centrifuged
complex food systems. Firstly, we identified the individual components and transferred to a 3-mm NMR glass tube.
of food polysaccharides mixtures from the 1H NMR spectra by means of
spectral line shape fitting (van Velzen et al., 2015; Wishart, 2008), also 2.4. NMR experiments
referred to as targeted profiling (Jacobs, van Velzen, & Mihaleva, 2013;
Mercier, Lewis, Chang, Baker, & Wishart, 2011). In this approach the The NMR-experiments were recorded with a zg30 pulse sequence on
NMR spectrum of the polysaccharide mixture is considered as the sum a Bruker Avance III 600 MHz NMR spectrometer equipped with a 5-mm
of the individual spectra of the pure components. Secondly, we ex- cryo-probe. The internal temperature of the probe was set at 290 K. The
tended the qNMR procedure such that it utilises the monomeric com- size of the FID was 96k. In total, 32 scans were collected with a re-
position of the polysaccharides, rather than its native forms laxation time of 10 s and an acquisition time of 4 s. The 90° pulse length
(Tojo & Prado, 2003) for quantification. We used the Saeman hydrolysis (∼12.5 μs) and receiver gain were determined automatically. The data
procedure (Carvalho de Souza et al., 2013; Mittal et al., 2009; Saeman was processed with Bruker TopSpin 3.2 software. Fourier transforma-
et al., 1954) to obtain the monomeric composition from the poly- tion with exponential window function and line broadening factor of
saccharides. Thirdly, we quantified the monomeric composition by 0.3 Hz were applied, followed by automatic phase correction and
means of Quantum Mechanical Total Line Shape (QMTLS) fitting baseline correction.
(Mihaleva et al., 2014; Tiainen, Soininen, & Laatikainen, 2014), and
furthermore adapted the qNMR method for polysaccharide-specific 2.5. QMTLS fitting
differences in (polymeric) dissolution and (monomeric) degradation
properties as a result of the hydrolysis. Finally, we used the estimated QMTLS fitting was performed with PERCH NMR software (PERCH
monosaccharide concentrations for quantification of polysaccharide solutions Ltd., Kuopui, Finland). A tailored spectral library was created
mixtures. This analysis requires prior knowledge on the polysaccharide for the line shape fitting of the monosaccharides of interest: glucose,
system and the associated monosaccharide stoichiometries. The per- galactose, mannose, rhamnose, xylose, fucose, arabinose, glucuronic
formance of the mixture analysis was tested with representative ternary acid and galacturonic acid. Each of the compounds was described by a
polysaccharide mixtures. set of spin particles and average, initial starting values for chemical
shifts, coupling constants, line widths and intensities (populations).
2. Materials and methods These parameters were obtained by fitting the pure model compounds
recorded in 13% (w/w) D2SO4 into the experimental spectra using the
2.1. Polysaccharide isolation and identification PERCH NMR software. All parameters were iteratively optimised. The
internal standard (maleic acid) was used to calculate the absolute
The full procedure for isolation and subsequent identification of the compound concentrations from the fitted populations. Python scripts
polysaccharide fraction from a food product matrix is described in were used for performing the QMTLS fitting in a non-supervised auto-
previous research (Grün et al., 2015). mated manner.

2.2. Solvents and reagents 2.6. Monosaccharide quantification

Deuterated water (D2O) was purchased from Euriso-Top (Saint Based on the NMR signal integral of the monosaccharide (Ix) and the
Aubin, France), deuterated sulphuric acid (D2SO4) was obtained from internal standard (IIS), the molar quantity (Nx) of the monosaccharide x
Sigma-Aldrich (Zwijndrecht, the Netherlands). All monosaccharides in the (reference) sample was estimated according to the general qNMR
(glucose, galactose, fucose, xylose, arabinose, rhamnose, mannose, equation (Malz & Jancke, 2005) (Eq. (1)):
glucuronic acid and galacturonic acid) were obtained from Sigma-
Ix ⋅NPIS⋅WIS⋅FIS⋅PIS
Aldrich. Xanthan gum (FNCS) was obtained from Jungbunzlauer Suisse Nx =
IIS⋅NPx⋅MWIS⋅VIS (1)
AG (Basel, Switzerland). GRINSTED® pectins and locust bean gum were
purchased from Danisco (Copenhagen, Denmark). The GRINSTED® where, N is the molar quantity (mol), I is the signal integral, NP is the
pectins had varying Degrees of Esterification (DE) and Degrees of number of protons, MW is the molecular weight (g/mol), W is the
Amidation (DA): LC810 (DE 37%), LC710 (DE 48%), LA210 (DE 34%, weight (g), V is the weight of the solvent volume (g), F is the volume
DA 17%). Rhamnogalacturonan (from soy bean pectin) from Megazyme fraction in the test solution (g) and P is the content mass fraction of the
(Bray, Ireland), tara gum (HV grade) from Kalys S.A. (Bernin, France), internal standard. Index x stands for the respective monosaccharide and
low acyl gellan (Kelcogel), high acyl gellan (Kelcogel LT100) and pectin IS for the internal standard.
LM5 (DE 10%) from CPKelco (Atlanta, USA) and guar gum and cellu- During the hydrolysis phase in the experiment, the monosaccharides
lose from Sigma-Aldrich. The 72% (w/w) D2SO4 in D2O solution was underwent degradation. The degree of degradation is dependent on
freshly prepared. Maleic acid (standard for qNMR, TraceCERT®), pur- monosaccharide x and the hydrolysis time th, and was expressed as the
chased from Sigma-Aldrich, was used as internal standard in D2O degradation factor Dx (th) . Dx (th) was estimated from the mono-
(0.5 mg/mL), and accurately weighed (0.01 mg) on an analytical bal- saccharide quantity in a Sugar Recovery Standard (SRS) before hydro-
ance. lysis (Nx (t 0) ) and after hydrolysis (Nx (th) ) following Eq. (2) (Carvalho de

380
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

Souza et al., 2013): 2.8. Validation of polysaccharide quantification


Nx (th)
Dx (th) = 1 − = k x th Repeatability (r) and the within-laboratory reproducibility (R) were
Nx (t 0) (2)
determined on a series of twelve different food polysaccharides. Both r
where, kx is the rate constant (min−1) of the degradation. Eq. (2) was and R were pooled over three concentrations levels (5, 10 and 15 mg/
obtained by solving the differential equation of a zero-order reaction. mL) that were measured in duplicate on four different days. Moreover,
Here we assumed a constant degradation rate of x during the hydrolysis for a commercially available microcrystalline cellulose (Sigma Aldrich)
phase. the trueness, linear working range and detection limit were determined.
When quantifying monosaccharides using Eq. (1), and estimating The minimum specified purity of 95% (w/w), as given by the supplier,
the associated degradation factor using Eq. (2), the general convention was accepted as the reference value. Trueness was estimated with blank
is that all dissolved monosaccharides are present at t0. This assumption samples that were spiked with increasing amounts of the cellulose
holds for the SRS solution, but does not hold for polysaccharide systems standard within the concentration range (0–60 mg/mL). The cellulose
where monosaccharides are gradually released during hydrolysis. To standard was dried for 24 h at 70 °C in a drying oven prior to the
prevent overestimation of the results, a modification of the degradation analysis. The linear working range was estimated for the same con-
factor Dx (th) was required. centration range. Estimated cellulose concentrations were compared
The release of monosaccharide x at timepoint tr, which is dependent with the actual cellulose concentrations using a Lack-of-Fit test. The
on the polysaccharide y, is denoted as the derivative function f ′ y (t r) . limit of quantification was based on an estimate of (10 times) the
The associated non-derivative function f y (t r) represents the cumulative standard deviation from 10 calibration samples containing a small
dissolution curve of y. This dissolution function f y (t r) was directly es- amount of the cellulose standard (1 mg).
timated from a pure polysaccharide standard with a series of NMR
experiments collected over a time interval [to, th]. A selected char- 2.9. Polysaccharide mixture analysis
acteristic isolated polysaccharide peak was (semi)-quantified on each
measured timepoint by manual integration (Fig. 3). When combining Quantitative analysis of polysaccharide mixtures needs additional
the release function ( f ′ y (t r) ) and the rate constant of degradation (kx), information on the molar monomeric stoichiometry (per polysaccharide
the monosaccharide quantity Nxy (t r) was calculated at any timepoint tr y), in addition to Nx. The concatenated stoichiometries of all poly-
according to Eq. (3) using a solver (ODE45 in Matlab R2016a, The saccharides together form the stoichiometry matrix ST. In addition, if
MathWorks, Inc., Massachusetts, USA): the mixture contains J polysaccharides, Ny can be replaced by a row
tr vector Ny. The general quantitative equation for solving polysaccharide
Nxy (t r) = ∫ f ′ y (t r)(1 − k x (th − t r)) dt r mixtures (via least squares regression) is given in Eq. (8):
t0 (3)
Ny = (STS)−1Nx S (8)
in case a monosaccharide x does not degrade during hydrolysis, thus kx
Where, Ny is a row vector (1 × J) of molar polysaccharide quantities
is 0, then Nxy (t r) becomes solely dependent on the release function (Eq.
(mol), ST is a stoichiometry matrix (J × L), and Nx is a row vector
(4)). Here we introduce N 0y (t r) as the monosaccharide quantity (mol)
(1 × L) of molar monosaccharide quantities.
that could be obtained in the absence of any monosaccharide de-
To solve this regression problem, we applied an alternating least
gradation:
squares algorithm (PLS Toolbox 7.8.2, Eigenvector Research, Inc.) with
tr
non-negativity constraints on Ny and equality constraints on ST in
N 0y (t r) = ∫ f ′ y (t r) dt r Matlab R2016a.
t0 (4)

similar to Eq. (2), the modified degradation factor Dxy (t r) was calcu- 2.10. Validation of polysaccharide quantification
lated (Eq. (5)). Note that Dxy (t r) is dependent on x and y:
Nxy (t r) A series of polysaccharide mixtures were analysed according to a
Dxy (t r) = 1 − full 33 three-level factorial design with a centre point. Three poly-
N 0y (t r) (5) saccharides (factors) were included in the design, i.e. high acyl gellan,
the monosaccharide concentration (Nx (t r) , mol) was finally calculated xanthan and LBG. The estimated polysaccharide ratios were compared
with the modified degradation factor according to Eq. (6): with the actual (theoretical) quantities in order to investigate the ac-
curacy of the mixture analysis. All 9 combinations of the three poly-
Ix ⋅NPIS⋅WIS⋅FIS⋅PIS 1
Nx (t r) = ⋅ saccharides were investigated in 5-fold at three different concentrations
IIS⋅NPx⋅MWIS⋅VIS 1 − Dxy (t r) (6) levels (5 mg, 2.5 mg, 0 mg). Agreement between estimated and theo-
retical values was evaluated in a trueness study and tested on off-set
bias and proportional bias.
2.7. Polysaccharide quantification

Quantification of pure polysaccharides can be directly achieved 3. Results and discussion


from the estimated concentration levels of the individual mono-
saccharides (Eq. (6)). When considering a collection of L mono- 3.1. Identification of polysaccharides in food systems
saccharides for a polysaccharide y, we can define a row vector Nx
(1 × L) containing all molar monosaccharide quantities. Multiplication The first step towards the quantification of polysaccharides in food
of Nx with their associated molecular weights (MWx, size (L × 1)) re- systems is their isolation and subsequent identification. Recently, an
sults in the absolute total weight of that particular polysaccharide (Wy extensive isolation procedure for unbiased and full recovery of poly-
in g) in the sample material as denoted in the matrix calculation in Eq. saccharides from complex food matrices has been described (Grün
(7). Note that the molecular weights of ‘anhydrous’ monomeric units in et al., 2015). This procedure yields isolates of polysaccharides that are
polysaccharides differ from free ‘hydrous’ monosaccharides (Δ = 18 g/ devoid of low molecular weight molecules. Fig. 1 shows an example of a
1
mol). H NMR spectrum of a polysaccharide isolate that was obtained from a
food product, where the individual polysaccharides were present at 1%
Wy = Nx MWx (7) (w/w) or less. Even though the signals are broad and show a large

381
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

Fig. 1. 1H NMR spectrum (600 MHz) of a bouillon


cube isolate ( ). The polysaccharides in this
isolate were identified by line shape fitting using the
pure spectra of guar gum ( ), LBG ( ), low
acyl gellan ( ) and xanthan ( ).

degree of overlap, the spectral contributions of different poly- them for food-specific polysaccharides. Firstly, for xanthan and pectin
saccharides could clearly be recognised. Component identification is the hydrolysis time was extended to 180 min (instead of 90 min) to
aided by fitting the 1H NMR spectra of pure polysaccharides into the 1H attain full conversion of the polysaccharides into their monosaccharide
NMR spectrum of the isolate. This approach is referred to as line shape constituents. To unify the hydrolysis protocol for all polysaccharides,
fitting or curve fitting (Wishart, 2008) and showed that the mixture 1H this prolonged hydrolysis time was standardised in the method. Sec-
NMR spectrum was largely explained by the sum of few poly- ondly, we introduced an adapted pre-solubilisation step at room tem-
saccharides. Fig. 1 showed that the mixture analysis lead to the un- perature without temperature control. This makes the method faster
ambiguous identification of four different polysaccharides which are and simpler when compared with the original protocol that aimed for a
commonly found in food (guar gum, xanthan, LBG and low acyl gellan). temperature-controlled pre-solubilisation in a thermostatic water bath
However, due to the unknown purity of most polysaccharide standards (30 °C). This adaptation does not affect the method performance and
we deem the line shape fitting approach unsuitable for direct quanti- makes it more convenient for high-throughput purposes. In the data
fication. Another downside of this method is that certain poly- acquisition, a few optimisations were introduced as well. To avoid
saccharides (like xanthan) yield low-intense, featureless 1H NMR suppression of the anomeric sugar-signals (δ 3.7–4.6 ppm), 1H NMR
spectra which can hardly be distinguished from other polysaccharides spectra were recorded with a 30° pulse sequence without water sup-
in complex mixtures. In addition, subtle differences in spectral features pression. In addition, the relaxation delay was set to 10 s, thereby en-
between comparable polysaccharides within polysaccharide families hancing the analysis throughput (up to six samples per hour) while
(e.g. galactomannans) may not be sufficient for attaining unambiguous maintaining the quantitative nature of the 1H NMR spectra. High field
identifications in all cases. Therefore, we proceeded with a high- NMR was required to obtain the desired spectral resolution on the α-
throughput approach in which the polysaccharides (or mixtures of and β-anomeric monosaccharide signals, which were used for the
polysaccharides) are converted into their constituent monosaccharides quantification. For quantification, we used both the α- and β-anomeric
prior to the quantification (Carvalho de Souza, 2017). monosaccharide signals. Up to nine different types of monosaccharides
were observed in the final hydrolysates. In mixtures, these mono-
saccharides often have overlapping anomeric signals, which hinder
3.2. Rapid Saeman hydrolysis on food polysaccharides manual quantification (Fig. 2). Quantification based on Quantum Me-
chanical Total Line Shape (QMTLS) fitting offers a suitable solution in
The Saeman procedure was used to hydrolyse the polysaccharides mixture analysis (Tiainen et al., 2014). QMTLS fitting can be used for
into their monomeric units. With 1H NMR it is possible to analyse and estimating the signal area of overlapping anomeric doublets in-
quantify the monosaccharides in the Saeman hydrolysates without ex- dependently and has a major advantage that the calculations can be
tensive additional sample preparation (Carvalho de Souza et al., 2013; executed in automation.
Mittal et al., 2009). We adapted the developed methods and optimised

Fig. 2. 1H NMR spectrum (600 MHz) of the anomeric


region of the nine monosaccharides considered in
this study: glucose (Glc), galactose (Gal), arabinose
(Ara), fucose (Fuc), xylose (Xyl) mannose (Man),
rhamnose (Rha), glucuronic acid (GlcA) and ga-
lacturonic acid (GalA).

382
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

Fig. 3. 1H NMR stacked spectra (600 MHz) of the


hydrolysis of a pectin, sampled at 15 min intervals
for the first 90 min, and sampled at 30 min intervals
for the last 90 min. The greyed area (δ
4.42–4.18 ppm), annotated to eCHOH-signals of the
pectin backbone, is used to estimate polysaccharide
degradation.

3.3. Non-instantaneous polysaccharide hydrolysis progressing degradation can be monitored. In Fig. 4A, the time courses
of both pectin degradation and monosaccharide release were plotted,
While optimising the method, we detected a systematic bias be- thereby indicating that the disappearance of the polysaccharide is in-
tween the measured quantities and the true quantities for particular versely related to the release of monosaccharides. In Fig. 4A we ob-
polysaccharides. The measured quantities for pectin for example were served that monomeric release is non-instantaneous. Instead, we could
overall larger than their accepted values. It is however well known that identify three consecutive stages from the experimental data. In the
during the Saeman procedure (in strong acidic conditions at elevated initial lag phase of hydrolysis (0–1 h), pectin partially hydrolyses into
temperatures) not only hydrolysis of the polysaccharides takes place oligomers (Coenen, Bakx, Verhoef, Schols, & Voragen, 2007). As the
but also degradation of the released monosaccharides (Shi, Yokoyama, molecular mobility of pectin increased, also their associated 1H NMR
Akiyama, Yashiro, & Matsumoto, 2012). Hence current monosaccharide signals in Fig. 3 sharpen up. Degradation to monosaccharides particu-
quantification methods use a degradation factor to correct for the larly occurred in the second hour. In this second phase, the majority of
monosaccharide degradation (Eq. (2)). A major advantage of 1H NMR all starting material (90%) degrades. After two hours, hydrolysis
in combination with Saeman monosaccharide analysis is that poly- reaches its decelerating phase. In this final stage, the remaining 10% of
saccharides and monosaccharides can be observed simultaneously, thus pectin is converted at lower reaction rate. The characteristics of the
allowing for direct assessment of the hydrolysis state. The time course concentration-time curves in Fig. 4A are pectin-specific. Different
of hydrolysis is demonstrated in Fig. 3 over a period of 180 min for the polysaccharides show different trends (Fig. 4B). For example, the de-
polysaccharide pectin. From top to bottom the 1H NMR spectra showed gradation of cellulose and xanthan is rapid and linear, whereas pectin,
increasing signals of released monosaccharides (e.g. δ 4.55 ppm, δ as shown before, typically demonstrates a slow, non-continuous de-
3.97 ppm, δ 3.86 ppm, δ 3.65 ppm, δ 3.56 ppm, and δ 3.50 ppm) and at gradation behaviour.
the same time the disappearance of pectin (e.g. δ 4.42–4.18 ppm and δ The concentration-time characteristics during hydrolysis were es-
3.75–3.66 ppm). When integrating the characteristic signal area of tablished for a range of polysaccharides that are commonly used in food
pectin within the spectral region δ 4.42–4.18 ppm (eCHOH), the products. We found that the conversion into monosaccharides could be

Fig. 4. A) Time course of the pectin degradation


( ), which is inversely related to the time course
of monosaccharide release ( ). The actual mea-
surement points include error bars, indicating the
inaccuracy in the individual estimations (standard
deviation). B) Monosaccharide release functions for
the polysaccharides pectin LC810 ( ), xanthan
( ), and cellulose ( ).

383
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

Table 1
Overview of the Dx and Dxy-values and calculation and validation results for a set of twelve polysaccharides.

Polysaccharide (y) Monosaccharide (x) Dx Dxy Concentration x based on Dx Concentration x based on Dxy Overestimation of Dx RSDra RSDRb
% (w/w) % (w/w) (%) (%) (%)

Cellulose Glc 0.07 0.07 92 92 0 1.8 3.3


Xanthan Glc 0.07 0.06 28 28 1 2.1 3.9
Man 0.10 0.09 20 20 1 3.7 4.3
GlcA 0.19 0.06 12 12 1 6.6 8.0
Locust Bean Gum Gal 0.10 0.07 16 16 1 4.5 8.7
Man 0.10 0.09 58 57 1 2.8 5.9
Guar Gum Gal 0.10 n.a.c 23 – – 4.0 6.7
Man 0.10 n.a.c 40 – – 2.9 3.4
Tara Gum Gal 0.10 0.07 18 18 1 3.7 10.8
Man 0.10 0.09 59 58 1 2.7 5.3
High Acyl Gellan Glc 0.07 0.06 25 24 1 3.9 5.0
Rha 0.09 0.08 10 10 1 4.6 5.3
GlcA 0.19 0.06 5 5 1 9.0 13.3
Low Acyl Gellan Glc 0.07 0.06 30 29 1 3.1 5.2
Rha 0.09 0.08 16 16 1 3.3 4.6
GlcA 0.19 0.06 7 7 1 12.6 18.8
Pectin LC 810 GalA 0.30 0.16 54 45 20 3.0 3.1
Pectin LC 710 GalA 0.30 0.16 53 44 20 5.0 7.8
Pectin LA 210 GalA 0.30 0.16 54 45 20 3.5 4.0
Pectin LM 5 GalA 0.30 0.16 40 34 20 5.1 5.6
Rhamnogalacturonan GalA 0.30 n.a.c 37 – – 5.6 7.7
Rha 0.09 n.a.c 7 – – 2.5 3.2

a
RSDr: Relative standard deviation of repeatability, based on calculations with Dx.
b
RSDR: Relative standard deviation of within-laboratory reproducibility, based on calculations with Dx.
c
n.a.: not analysed.

well-described via a polysaccharide-specific cumulative release func- trueness, we also determined the linear working range and the limit of
tion (fy). Moreover, it does not occur instantaneously at to. Among the quantification. Based on a Lack-of-Fit test (F < Fcritical), the linearity of
investigated polysaccharides, the release function for pectin was most the method was tested and proved to be significant over the in-
distinctive, due to its relatively long lag phase prior to monosaccharide vestigated range (0–60 mg/mL). Based on the limit of detection, we
release. Because of this long lag phase, the effective time that remains determined that a minimum of 0.24 mg cellulose was required for at-
for monosaccharide degradation was relatively short. This should be taining recovery estimates with a relative standard deviation of 10%.
accounted for in the quantification when determining the degradation
factor Dx (Eq. (2)). Since Dx is based on the instantaneous release of
3.5. Quantification of polysaccharide content
monosaccharides at to, and does not consider fy, the end results will be
prone to overestimation. We therefore aimed for an adapted version of
An essential part for the quantification of polysaccharide mixtures
Dx, defined as Dxy (Eq. (5)) to compensate for delayed monosaccharide
was the integration of prior knowledge on the individual poly-
release and resultant estimation bias. For a wide range of poly-
saccharides. Hence, for each polysaccharide, a reliable estimate of the
saccharides, the Dx and Dxy values and their corresponding mono-
molar monomeric stoichiometry was needed. The molar stoichiometry
saccharide concentrations were determined (Table 1). Especially for
for each polysaccharide could directly be derived from its theoretical
pectin, the overestimation of Dx is significant (20%), whereas for i.e.
structure. In the current experiment, however, the molar stoichiometry
xanthan this overestimation is insignificant (1%). As expected, the
was empirically estimated from polysaccharide reference standards.
importance of Dxy increases with the instability of monosaccharides (i.e.
We demonstrate the applicability of the mixture analysis by com-
GalA and Xyl) and the stability of the polysaccharide (pectin) under
posing binary and ternary mixtures of polysaccharides that commonly
hydrolysis conditions.
occur in food products, i.e. high acyl gellan, xanthan and locust bean
gum (LBG). This experiment was designed according to a 33 three-level
3.4. Quantification and validation of pure polysaccharide standards factorial design with a centre point. Signal integration in the 1H NMR
spectra of the resulting hydrolysates was done manually as well as by
With NMR, we could identify and quantify the monomeric content QMTLS fitting. The two different approaches each resulted in a separate
(%w/w) of a series of pure reference polysaccharides. The precision of matrix of monosaccharide quantities (Nx). Based on the stoichiometry
the estimations (in terms of r and R) were determined and presented in matrix ST the individual polysaccharide quantities (Ny) were estimated
Table 1. Overall, the method resulted in a good precision in terms of via a non-negative least squares regression (Eq. (8)). The estimated
relative standard deviation for repeatability (RSDr = 4.1 ± 0.9%) and mixture ratios that were obtained from manual signal integration and
within-laboratory reproducibility (RSDR = 6.1 ± 1.4%). QMTLS fitting were plotted in the ternary diagrams in Fig. 5A and B,
In theory, the sum of the individual monosaccharide levels per respectively. The accuracy of the QMTLS-based procedure was slightly
polysaccharide in Table 1 should provide an estimate of the total better than the manual approach as demonstrated by the relative
polysaccharide recovery. However, since the exact purities of these tighter scatter patterns of the estimated values around the design
polysaccharides were unknown, or not well specified by the suppliers, points.
the estimated polysaccharide concentrations could not be compared For both quantification methods, the trueness was established as no
against their true values. Only for cellulose, we performed a trueness significant off-set and proportional bias were detected between the
study because the product purity was specified. The recovery for cel- estimated and the theoretical values. This observation indicated that
lulose was 92%, and neither a proportional nor an off-set bias were the non-negative least squares regression approach, with ST as prior
detected when compared to the true value (95% w/w). This indicates knowledge, is a valid quantitative approach to resolve polysaccharide
that the method was not only precise, but also accurate. Besides mixtures. Correct molar monosaccharide stoichiometries are therefore

384
D.W.H. Merkx et al. Carbohydrate Polymers 179 (2018) 379–385

Fig. 5. Ternary diagrams representing the predicted


mixture ratios of xanthan, LBG and high acyl gellan.
Quantification of the three polysaccharides from the
1
H NMR spectra was based on (A) manual peak in-
tegration, and (B) QMTLS fitting.

critical for accurate polysaccharide quantifications. Carvalho de Souza, A. (2017). Quantification of food polysaccharides by means of NMR.
In G. A. Webb (Ed.). Modern magnetic resonance (pp. 1–19). Springer International
Publishing.
4. Conclusions Coenen, G. J., Bakx, E. J., Verhoef, R. P., Schols, H. A., & Voragen, A. G. J. (2007).
Identification of the connecting linkage between homo- or xylogalacturonan and
rhamnogalacturonan type I. Carbohydrate Polymers, 70(2), 224–235.
A method has been presented for the quantitative assessment of De Ruiter, G. A., Schols, H. A., Voragen, A. G. J., & Rombouts, F. M. (1992). Carbohydrate
pure polysaccharides and polysaccharide mixtures. The Saeman NMR analysis of water-soluble uranic acid-containing polysaccharides with high-perfor-
method was tailored towards a convenient method for a wide range of mance chromatography using methanolysis combined with TFA hydrolysis is superior
to four other methods. Analytical Biochemistry, 185, 176–185.
food polysaccharide systems. The method was repeatable Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the properties of
(RSDr = 4.1 ± 0.9%) and reproducible (RSDR = 6.1 ± 1.4%) with a dispersed systems. Food Hydrocolloids, 17(1), 25–39.
high throughput (40–60 samples a day). We employed automated Englyst, H. N., & Cummings, J. H. (1984). Simplified method for the measurement of total
non-starch polysaccharides by gas – liquid chromatography of constituent sugars as
mixture analysis according to a non-negative least squares analysis in
alditol acetates. Analyst, 109(7), 937–942.
order to increase the spectral processing speed and decrease operator Grün, C. H., Sanders, P., van der Burg, M., Schuurbiers, E., van Adrichem, L., van Velzen,
errors. This mixture analysis required a priori knowledge on the iden- E. J., ... Schols, H. A. (2015). Strategy to identify and quantify polysaccharide gums in
tity of the individual polysaccharides as well as on the molar stoi- gelled food concentrates. Food Chemistry, 166, 42–49.
Jacobs, D. M., van Velzen, E. J. J., & Mihaleva, V. (2013). Evaluation of approaches for
chiometries of the monosaccharide constituents. For the mono- quantitative targeted profiling of complex compositions using 1D 1H NMR spectroscopy.
saccharide quantification, an adapted degradation factor was Magnetic resonance in food science: Food for thought3–13.
implemented which accounts for both the monomeric degradation and Malz, F., & Jancke, H. (2005). Validation of quantitative NMR. Journal of Pharmaceutical
and Biomedical Analysis, 38(5), 813–823.
the non-instantaneous polysaccharide degradation. Mercier, P., Lewis, M. J., Chang, D., Baker, D., & Wishart, D. S. (2011). Towards auto-
The entire strategy for identifying and quantifying polysaccharides matic metabolomic profiling of high-resolution one-dimensional proton NMR spectra.
and polysaccharide mixtures provides a fast and accurate analytical Journal of Biomolecular NMR, 49(3), 307–323.
Mihaleva, V. V., Korhonen, S.-P., van Duynhoven, J., Niemitz, M., Vervoort, J., & Jacobs,
tool. The combined method demonstrated its potential for a wide ap- D. M. (2014). Automated quantum mechanical total line shape fitting model for
plication range within food analysis. quantitative NMR-based profiling of human serum metabolites. Analytical and
Bioanalytical Chemistry, 406(13), 3091–3102.
Mittal, A., Scott, G. M., Amidon, T. E., Kiemle, D. J., & Stipanovic, A. J. (2009).
Acknowledgements Quantitative analysis of sugars in wood hydrolyzates with 1H NMR during the au-
tohydrolysis of hardwoods. Bioresource Technology, 100(24), 6398–6406.
We gratefully acknowledge the expert advice from Adriana Saeman, J. F., Moore, W. E., Mitchell, R. L., & Millet, M. (1954). Techniques for the
determination of pulp constituents by quantitative paper chromatography. Tappi, 37,
Carvalho de Souza (DSM, Delft) and Christian Grün (Unilever R & D
336–343.
Vlaardingen). We thank Seenakshi Dauwan (Unilever R & D Saha, D., & Bhattacharya, S. (2010). Hydrocolloids as thickening and gelling agents in
Vlaardingen) for carefully performing NMR measurements. We thank food: A critical review. Journal of Food Science and Technology, 47(6), 587–597.
Velitchka Mihaleva (Unilever R & D Vlaardingen) for writing the core of Shi, Y., Yokoyama, T., Akiyama, T., Yashiro, M., & Matsumoto, Y. (2012). Degradation
kinetics of monosaccharides in hydrochloric, sulfuric, and sulfurous acid.
the Python scripts. BioResources, 7(3), 4085–4097.
Tiainen, M., Soininen, P., & Laatikainen, R. (2014). Quantitative quantum mechanical
Appendix A. Supplementary data spectral analysis (qQMSA) of (1)H NMR spectra of complex mixtures and biofluids.
Journal of Magnetic Resonance, 242, 67–78.
Tojo, E., & Prado, J. (2003). A simple 1H NMR method for the quantification of carra-
Supplementary data associated with this article can be found, in the geenans in blends. Carbohydrate Polymers, 53(3), 325–329.
online version, at http://dx.doi.org/10.1016/j.carbpol.2017.09.074. van Duynhoven, J., van Velzen, E., & Jacobs, D. M. (2013). Quantification of complex
mixtures by NMR. Annual reports on NMR spectroscopy181–236.
van Velzen, E. J. J., Dauwan, S., de Roo, N., Grun, C. H., Westphal, Y., & van Duynhoven,
References J. P. M. (2015). Quantitative NMR assessment of polysaccharides in complex food ma-
trices. Magnetic resonance in food science: Defining food by magnetic resonance. The
Royal Society of Chemistry39–48.
Ahmed, A. E. R., & Labavitch, J. M. (1977). A simplified method for accurate determi-
Wang, S., Li, C., Copeland, L., Niu, Q., & Wang, S. (2015). Starch retrogradation: A
nation of cell wall uronide content. Journal of Food Biochemistry, 1, 361–365.
comprehensive review. Comprehensive Reviews in Food Science and Food Safety, 14(5),
Bertaud, F., Sundberg, A., & Holmbom, B. (2002). Evaluation of acid methanolysis for
568–585.
analysis of wood hemicelluloses and pectins. Carbohydrate Polymers, 48, 319–324.
Wishart, D. S. (2008). Quantitative metabolomics using NMR. TrAC Trends in Analytical
Carvalho de Souza, A., Rietkerk, T., Selin, C. G., & Lankhorst, P. P. (2013). A robust and
Chemistry, 228–237.
universal NMR method for the compositional analysis of polysaccharides.
Carbohydrate Polymers, 95(2), 657–663.

385

You might also like