A method of water-soluble solid fraction saturation concentration evaluation in dry thalli of Antarctic lichenized fungi, in vivo.
Journal: 2017/September - Biochemistry and Biophysics Reports
ISSN: 2405-5808
Abstract:
BACKGROUND
At initial steps of rehydration from cryptobiosis of anhydrobiotic organisms or at rehydration of dry tissues the liquid 1H NMR signal increased anomaly. The surplus in liquid signal may appear if some solid constituents dissolved, or if they were decomposed by enzymatic action.
METHODS
Hydration kinetics, sorption isotherm, 1H NMR spectra and high power relaxometry were applied to monitor gaseous phase rehydration of Antarctic lichen Cetraria aculeata. Tightly and loosely bound water signal were distinguished, and the upper hydration limit for dissolution of water soluble solid fraction was not observed. A simple theoretical model was proposed.
RESULTS
The hydration courses showed a very tightly bound water fraction, a tightly bound water, and a loosely bound water fraction. Sigmoidal in form sorption isotherm was fitted well by multilayer sorption model. 1H NMR showed one Gaussian signal component from solid matrix of thallus and one or two Lorentzian line components from tightly bound, and from loosely bound water. The hydration dependency of liquid signal was fitted by rational function.
CONCLUSIONS
Although in dehydrated C.aculeata the level of carbohydrates and polyols was low, the lichenase action during rehydration process increased it; the averaged saturation concentration cs =(57.3±12.0)%, which resembled that for sucrose.
CONCLUSIONS
The proposed method of water soluble solid fraction saturation concentration, cs , calculation from 1H NMR data may be applied for other organisms experiencing extreme dehydration or for dry tissues. We recalculated the published elsewhere data for horse chestnut (Aesculus hippocastanum) bast [water-soluble solid fraction recognized as sucrose, cs =(74.5±5.1)%]; and for Usnea antarctica, where cs =0.81±0.04.
Relations:
Content
Citations
(2)
References
(9)
Drugs
(2)
Affiliates
(2)
Similar articles
Articles by the same authors
Discussion board
Biochemistry and Biophysics Reports. Jun/30/2016; 6: 226-235
Published online Apr/21/2016

A method of water-soluble solid fraction saturation concentration evaluation in dry thalli of Antarctic lichenized fungi, in vivo

Abstract

Background

At initial steps of rehydration from cryptobiosis of anhydrobiotic organisms or at rehydration of dry tissues the liquid 1H NMR signal increased anomaly. The surplus in liquid signal may appear if some solid constituents dissolved, or if they were decomposed by enzymatic action.

Methods

Hydration kinetics, sorption isotherm, 1H NMR spectra and high power relaxometry were applied to monitor gaseous phase rehydration of Antarctic lichen Cetraria aculeata. Tightly and loosely bound water signal were distinguished, and the upper hydration limit for dissolution of water soluble solid fraction was not observed. A simple theoretical model was proposed.

Results

The hydration courses showed a very tightly bound water fraction, a tightly bound water, and a loosely bound water fraction. Sigmoidal in form sorption isotherm was fitted well by multilayer sorption model. 1H NMR showed one Gaussian signal component from solid matrix of thallus and one or two Lorentzian line components from tightly bound, and from loosely bound water. The hydration dependency of liquid signal was fitted by rational function.

Conclusions

Although in dehydrated C.aculeata the level of carbohydrates and polyols was low, the lichenase action during rehydration process increased it; the averaged saturation concentration cs=(57.3±12.0)%, which resembled that for sucrose.

General significance

The proposed method of water soluble solid fraction saturation concentration, cs, calculation from 1H NMR data may be applied for other organisms experiencing extreme dehydration or for dry tissues. We recalculated the published elsewhere data for horse chestnut (Aesculus hippocastanum) bast [water-soluble solid fraction recognized as sucrose, cs=(74.5±5.1)%]; and for Usnea antarctica, where cs=0.81±0.04.

Graphical abstract

The combined fit of liquid to solid component of 1H NMR signal, L/S, and respective fit of immobilized water signal to total liquid signal, L1/L, expressed as a sample hydration level, Δm/m0, allows to get saturation concentration of water-soluble solid fraction, cs, in extremely dry biological system (cryptobiotic organism, in vivo).

fx1

Highlights

  • Hydration kinetics show three fractions of bound water in lichen thallus.

  • Sorption isotherm yields water fraction saturating primary water binding sites.

  • 1H NMR spectra distinguish solid thallus and mobile and immobilized water fractions.

  • Signal intensities supply saturation concentration, cs, of water-soluble solid fraction.

  • A simple theoretical model is proposed and applied also to already published NMR data.

1Introduction

At initial steps of a very mild rehydration of some dry biological systems the liquid component of 1H NMR signal does not increase proportionally with the increasing hydration level. Usually the observed dependency is well fitted using the rational function, e.g. in cellular organelles such as freeze-dried photosynthetic membranes [1], [2]; in plant tissues like horse chestnut bast [3]; at the initial phases of seed imbibition [4], as well as in whole living organisms, e.g. in thalli of several species of lichenized fungi [5], [6], [7], [8], [9].

In extremely dry biological systems the carbohydrates and polyols remain completely/partially solidified (actual water concentration is far below their saturation hydration concentration point) and gradually dissolve with the increasing hydration level.

Carbohydrates and/or polyols play a critical role either for freezing resistivity [10] or in acute desiccation shock resistivity for numerous living organisms. For example, before dehydration to cryptobiosis Polypedilum vanderplanki larvae [11], [12] accumulate large quantities of trehalose [13], [14], [15], [16]. The major carbohydrate constituents of seed are starch (up to 80% dry mass of seed) and cellulose. In developing seed enzyme action decomposes starch to simple sugars, which remain in aqueous medium [4], [17], [18], [19]. The accumulation of monosaccharides and polyols in lichen mycobiont is an adaptive feature for dealing with the complete dehydration [20], [21]. Ribitol level modifies the photosynthetic activity in some green algal foliose lichens [22].

1H NMR relaxometry was applied to the analysis of extremely dry biological systems with a limited amount of water-soluble solid fraction. In horse chestnut bast it allowed to identify the water-soluble solid fraction as sucrose [3]. With the increased hydration level, as far as water-soluble solid fraction dissolves, the hydration dependency of total liquid 1H NMR signal expressed in units of solid, L/S, is described by rational function, whereas for higher hydration levels at which the whole water-soluble solid fraction is completely dissolved the L/S hydration dependency becomes linear [3]. The presence of a threshold hydration level allows one to calculate the saturation concentration of water-soluble solid fraction, which may be used to identification of such a fraction.

In this report we propose a method for evaluation of the water-soluble solid fraction saturation concentration in case of a high amount of water-soluble solid fraction if the threshold hydration level is not detected, and if in the investigated biological system the signal coming from tightly bound water, L1, and from loosely bound water signal, L2, may be distinguished. A simple 1H NMR experiment should be combined with sorption analysis for the investigated system.

As a model system we used a fruticose lichen Cetraria aculeata (Schreb.) Fr. thalli collected from sites in maritime Antarctica. Beside proton NMR spectra and relaxometry we recorded sorption kinetics, and sorption isotherms. The measurements were performed in vivo.

2Theory

Harańczyk, Węglarz and Sojka [3] proposed that if the sample is hydrated by the mass, Δm, of water and if the water soluble solid fraction is present in the system, the amplitude of FID for liquid/mobile signal component may be expressed as a sum of three terms:(1a)L=L0+αH2OρH2OΔm+αcdρcmcdwhere L0 (if such a fraction is detected) is the signal of water'sealed’ in pores of dry solid matrix [1], [2], [23], the second term is proportional to the mass of water added, and the last one comes from water soluble solid fraction already dissolved. The amplitude of FID for solid signal component is then given by:(1b)S=S0αcuρcmcd,and:(1c)S0=αsρsm0,where αH2O, αcd, αcu are the proportionality coefficients describing the effective contribution of a given proton pool to the total signal for water, and for water-soluble solid fraction in liquid and in solid phase, respectively (see Fig. 1). The values of coefficients α may be decreased e.g. by the presence of paramagnetic ions in the solution or on the surfaces of the solid matrix. If electron paramagnetism is absent in the investigated system αi=1[24]. However, in some biological systems solid NMR signal may be decreased by the presence of endogeneous electron paramagnetic ions, e.g. manganese [25]. The ρH2O and ρcare proton densities of water and of water-soluble solid fraction, respectively, whereas mcd is the mass of dissolved water-soluble solid fraction.

Fig. 1

Proton subsystems distinguished at rehydration of C. aculeata in dry thalli (a), at low hydration level (b), and in wet thalli (c): solid thallus, S0, (black space); air (white space); undissolved water soluble solid fraction (rectangular grid); immobilized liquid fraction consisted of water soluble solid fraction dissolved to saturation concentration, cs (dots), mobile liquid fraction consisted of water soluble solid fraction dissolved to cs, and mobile water sealed in pores of dry solid (L0) (fine dots).

Fig. 1.

Knowing that the saturation concentration cs of the water-soluble fraction is given by:(2)cs=mcdΔm+mcd.

The dissolved portion of water-soluble solid fraction, mcd, may be obtained from:(3)mcd=cs1csΔm.

In microheterogeneous systems bound water tend to orient depending on its mobility, and even tends to organize in some well defined layers [26]. If tightly bound water signal, L1, may be distinguished from loosely bound water fraction L2, the liquid signal component may be written as:(4)L=L0+L1+L2and:(5)Li=(αH2OiρH2O+αcdρccs1cs)Δmi,where denumerator i=1, 2, and Δm2=Δm−Δm1.

If the loosely bound (mobile) water pool filling the pores of the solid matrix is absent in the dry system, i.e. L0=0, and if the proportionality coefficients for mobile and for immobilized water are equal, αH2O1=αH2O2, and also for water soluble solid fraction is αcd=αcu, so, the L1 to L ratio may be written as:(6)L1L=m1m1+(1+γcs1cs)Δm2,where m1 is the mass of water saturating tightly bound water pool and the coefficient γ describes the water-soluble solid fraction proton density to water proton density ratio:(7)γ=ρcρH2O.

In the majority of living organisms which survive extreme dehydration the water-soluble solid fraction consists of sugars and/or polyols [5], [11], [12], [13], [14], [20]. Table 1 presents the saturation concentrations, and γ symbols for carbohydrates and polyols frequently occurring in biological systems. We used them to calculate the average values of γ. The coefficient γ does not vary much between biological carbohydrates with the averaged value equal γ=0.59±0.01. For polyols this value is quite similar and equals γ=0.67±0.06. If the type of water-soluble solid fraction is not known, the mean value averaged over sugars and polyols is equal γ=0.63±0.08.

Table 1

The saturation concentrations, cs[27], [28], [29], [30], [31], [32], proton densities and relative proton densities of carbohydrate and of the polyol fractions detected in photobiont and in mycobiont of Antarctic Cetraria aculeata thalli samples. For the carbohydrates the weighted average value of γ=0.59±0.01, whereas for polyols it equals γ=0.67±0.06.

Table 1
Sugarcsρcγ=ρcρH2O10−3·%
Fructose C6H12O60.440.0670.6010.203± 0.004
Glucose C6H12O60.540.0670.6010.085± 0.004
Sucrose C12H22O110.680.0640.5801.23 ±0.05
Trehalose C12H22O110.410.0640.5800.153 ±0.003
Polyolcsρcγ=ρcρH2O10−3·%
Arabitol C5H12O50.400.0790.71116.50±0.74
Myo-inositol C6H12O60.130.0670.6010.82 ±0.03
Mannitol C6H14O60.180.0770.6932.39 ±0.08

The solid signal component, S0, may be expressed as:(8)S0=αsρsms,where αs is the responsible proportionality coefficient for the solid matrix of the lichen thallus, ρs is the averaged proton density for the solid matrix of the thallus, and ms is the mass of the solid matrix of the thallus which in the absence of ‘sealed’ water fraction equals the dry mass of the sample, m0. If water-soluble solid fraction is present, we get for the solid signal, S:(9)S=αsρsm0αcuρcmcd.

Thus, the total signal from liquid component may be expressed in units of solid component, L/S, [from (5), (9)] as:(10)L/S(Δm/m0)=(k+αcdαsγcs1cs)Δm/m01αcuαsρcρscs1csΔm/m0.

Assuming that the decrease of the NMR signal for dissolved water-soluble solid fraction is comparable to this for water we get αcdαsk, and if the proportionality coefficient for the undissolved water-soluble the solid fraction is similar to the rest of the solid matrix of thallus, αcuαs1, we get:(11)L/S(Δm/m0)=k(1+γcs1cs)Δm/m01γδcs1csΔm/m0where:(12a)δ=ρsρH2O,and k is a slope of the L/S hydration dependence measured in case of the absence of water-soluble solid fraction expressed as:(12b)k=αLρLαSρS.

3Materials and methods

Cetraria aculeata (Schreb.) Fr thalli were collected on Penguin Island (South Shetlands Islands), Oceanic Antarctica (62°13′S, 58°25′W), in 2009 at the altitude 110 m above sea level. Air-dry thalli were stored in herbarium at relative humidity h≈40% and at room temperature with the hydration level, ∆m/m0=0.088±0.011, where m0 is the dry mass of the sample, and ∆m is mass of water taken up. Vitality tests performed using methylene blue staining showed that no less than 62% photobiont cells of thalli samples were alive before the measurements.

Polyol and sugar contents were measured using high resolution gas chromatography (HRGC) method. Dry lichen thalli were crushed in a mixed mill (MM 200, Retch, Germany) set at the vibrational frequency of 22 Hz for 2 min. Carbohydrates were extracted from 25 to 30 mg of pulverized thalli (in 1.5-ml tubes) with 800 μl of extraction mixture (methanol: chloroform: water, 12:5:3, v/v/v) containing 100 μg of d-chiro-inositol as an internal standard. The homogenate was shaken on Genie 2 vortex (Scientific Industries, USA) for 2 min, heated at 60 °C for 30 min and centrifuged at 20,000g for 20 min at 4 °C. Aliquots (400 μl) of clear methanol-water supernatant were desalted, evaporated to dryness. The dry residue was derivatized with a mixture of trimethylsilylimidazole: pyridine (1:1, v/v) as described previously [31]. The TMS-derivatives of carbohydrates were separated on a Zebron ZB-1 capillary column (Phenomenex, USA) in GC-2010 gas chromatograph (Shimadzu, Japan). The column was operated with an initial temperature 140 °C adjusted to 160 °C at 5 °C min−1, then to 200 °C at 10 °C min−1, and finally to 290 °C at 30 °C min−1 (with total resolution time 11 min). The injection port operated in the split mode (10:1) at 325 °C, and the flame ionization detector was maintained at 350 °C. Helium was used as a carrier gas (with linear velocity 40 cm s−1).

Soluble carbohydrates were quantified as described previously [33]. Standard of d-chiro-inositol (internal standard) was obtained from Industrial Research Ltd (New Zealand) and other carbohydrates were purchased from Sigma-Aldrich. The results of analyses are the means of three independent replicates±SE. Their values averaged for mycobiont and for photobiont are presented in Table 1.

Before the hydration courses thalli were incubated for 130 h over silica gel (h=0%), dehydrating to the hydration level ∆m/m0=0.031±0.003. The dehydration process was well fitted by a single exponential function with the dehydration time tD=(14.11±0.65) h (See Fig. 2).

Fig. 2

(a) The dehydration of air dry Cetraria aculeata (Schreb.) Fr. thalli to gaseous phase at the relative humidity h=0%, recorded as relative mass increase expressed in units of dry mass Δm/m0. (b) The residual function calculated as the difference between the fitted and recorded values of the relative mass increase; the reduced value of χ2=2.48.

Fig. 2.

The hydration time-courses were performed from the gaseous phase with controlled humidity, at room temperature (t=22 °C), over the surface of H3PO4 (h=9%), over the surface of saturated solutions of CH3COOK (h=23%), CaCl2 (h=32%), K2CO3 (h=44%), Na2Cr2O7 (h=52%), NH4NO3 (h=63%), Na2S2O3 (h=76%), K2CrO3 (h=88%), Na2 SO4 (h=93%), K2SO4 (h=97%), and over the water surface (h=100%). Hydration kinetics courses were performed on one given sample per each value of relative humidity. After completing of the hydration courses, the dry mass of the thallus was determined after heating at 70 °C for 72 h. Higher temperatures were not used as they may cause decomposition of some organic constituents of the thallus [34].

The air-dry thalli designed for the NMR measurements were chopped, placed in NMR tubes, and incubated over silica gel, then the hydration procedure was performed. The NMR measurements were performed on fourteen samples. The measurements of a drier sample covered a part of hydration range for the next more wet sample, whereas the hydration level was calculated as ∆m/m0.

1H NMR spectra were collected on Bruker Avance III spectrometer (Bruker Biospin), operating at the resonance frequency 300 MHz (at B0=7 T), with the transmitter power equal to 400 W. The pulse length was π/2=1.5 μs, bandwidths 300 kHz, and repetition time was 2 s.

1H NMR free induction decays (FIDs) were recorded using WNS HB-65 high power relaxometer (Waterloo NMR Spectrometers, St. Agatha, Ontario, Canada). The resonance frequency was 30 MHz (at B0=0.7 T); the transmitter power was 400 W; and the pulse length π/2=1.5 μs. In every FID 120 data points were collected, which were ordered in three equal groups with the different time-spacings equal to 0.4 μs, 2.5 μs, and 10 μs. The quadrature detection was applied, and the value of FID was a square root from the sum of squares of the phased and the dephased channel signals, Re2+Im2. FIDs were acquired using Compuscope 2000 card of an IBM clone controlling the spectrometer and averaged over 2000 accumulations.

The data obtained were analyzed using the FID analyzing procedure of a two-dimensional (in time domain) NMR signal-analyzing software CracSpin written at the Jagiellonian University, Cracow [35], or by commercially available software Origin 7.0.

4Results

4.1Hydration kinetics

The hydration courses for Cetraria aculeata were performed from the gaseous phase, and were fitted well by single exponential function for the relative humilities, h, controlled between 9% and 52% (see Fig. 3):(13a)Δm/m0=A0h+A1h[1exp(t/t1h)],where ∆m/m0 is the relative mass increase expressed in units of dry mass, m0, A1h is the saturation level for the fast component solely observed at a given relative humidity range, and A0h is the hydration level at h=0% (very tightly bound water fraction).

Fig. 3

The rehydration of the lichen Cetraria aculeata (Schreb.) Fr. from gaseous phase at different values of relative humidity h, recorded as relative mass increase expressed in units of dry mass Δm/m0. Targets humidity: h=9% – closed squares, h=23% – open squares, h=32% – closed circles, h=44% – open circles, h=52% – closed triangles, h=63% – open triangles, h=76% – closed reversed triangles, h=88% – open reversed triangles, h=97% – close starlets, h=100% – asterisks. The error bars are within the plot symbols.

Fig. 3.

As relative humidity exceeded 63% the slow hydration component appeared, and the hydration courses were fitted well by the two exponential function (Fig. 3):(13b)Δm/m0=A0h+A1h[1exp(t/t1h)]+A2h[1exp(t/t2h)],where A0h is the saturation level of very tightly bound water fraction (at h=0%), A1h and A2h are the saturation hydration levels for the fast and for the slow component, and t1h and t2h are hydration times for fast and slow component (tightly and loosely bound water fraction), respectively.

The averaged (over the all hydration courses) value of A0h=0.04±0.02. It is a very tightly bound fraction of water, which is not removed by dehydration over silica gel surface. Its relative mass is close to that obtained for other species of Antarctic lichens, e.g. for fruticose Usnea antarctica Du Rietz A0h=0.040±0.011, [6], and for foliose Umbilicaria aprina Nyl. A0h=0.054±0.011 [7].

The amplitude of the tightly bound water fraction equals to A1h=0.037±0.005, with the hydration time t1h=(0.43±0.10) h, whereas the loosely bound water fraction appearing for the courses performed from the values of relative humidity higher than 63% is characterized by the hydration time equal to t2h=(14.2±6.2) h. The saturation level of the loosely bound water component, A2h, increases gradually with increased humidity.

Loosely and tightly bound fractions may be distinguished by their proximity to the surface of solid matrix and thus by their mobility.

The total saturation hydration level, Ch=A0h+A1h+A2h, obtained at a given relative humidity, h, was taken for construction of sorption isotherm.

4.2Sorption isotherm

For Cetraria aculeata thalli the sorption isotherm is sigmoidal in form (Fig. 4), which usually is well fitted by the multilayer sorption models distinguishing two types of water binding sites on the surfaces of the investigated system, namely, (i) ‘primary’ water binding sites (directly bound to the adsorbent surface); and (ii) ‘secondary’, usually weaker, water binding sites (to the previously bound water molecules, or to the weaker binding sites on a solid matrix surface). The most commonly applied ones are the classic BET model [36], and GAB (Dent) model [37]. BET model takes a fixed value of the ratio of binding sites covered by n water molecules expressed in units of binding sites covered by n-1 water molecules b=Sn/Sn1|h=1=1, which is to some extent an artificial assumption, whereas in Dent model this number may be varied between 0 and 1 (somehow simulating the clustering effect), and is equal b=Sn/Sn1|h=1. Sorption isotherm for both models is expressed by:(14)Ch(h)=ΔMm0b1h(1bh)(1+b1hbh)where h is relative humidity expressed in absolute units, ΔM/m0 is the mass of water saturating primary binding sites, where Si is the number of binding sites covered by i water molecules, and the contribution of empty primary binding sites, S0, on the surface in units of sites with one water molecule, S1, at h=1 is expressed by the reciprocal of b1: S0/S1|h=1=1/b1.

Fig. 4

The sorption isotherm for Cetraria aculeata (Schreb.) Fr. The values of h represent the relative humidity, and the values of relative mass increase, ∆m/m0, are taken as the saturation values, Ch, from hydration kinetics (open circles = experimental data). Fitted Dent multilayer sorption model (Eq. (2)) (solid line = fitted Dent model), and ‘primary’ water binding sites monolayer coverage for Dent isotherm equal Δmm0=ΔMm0b1h1+(b1b)h[36] (dotted line = fitted ‘primary’ water binding sites population).

Fig. 4.

To test the relevance of the model used the sorption isotherm is usually presented in parabolic form (see Fig. 5), which for BET model is expressed as:(15a)hΔm/m0=A+Bh(A+B)h2,whereas for Dent model as:(15b)hΔm/m0=A+BhCh2,where parameters ΔMm0, b, b1 are connected with A, B, C by the formulas:(16a)b=B2+4ACB2A(16b)b1=BA+2b(16c)ΔMm0=1Ab1

Fig. 5

Parabolic form of Dent and BET model (open circles – experimental data, solid line – fitted Dent model (Eq. (3)b), dotted line – fitted BET model (Eq. (3)a)).

Fig. 5.

For Cetraria aculeata thalli the sorption isotherm is much better described by Dent model, as the parameter equals b=0.958±0.003. The sorption isotherm for other Antarctic lichens is also better fitted using Dent model, e.g. for Leptogium puberulum b=0.955 [8], for Umbilicaria aprina b=0.941 [7], and for Usnea antarctica b=0.913 [6].

The mass of water saturating primary water binding sites is equal ΔM/m0=0.046±0.003, which is very close to the value obtained from hydration kinetics for very tightly bound water pool.

The proportion of unoccupied water binding sites at h=1 equals for Cetraria aculeata 1/b1=0.925%. This result is comparable with values reported for other lichens, e.g. for U. aprina, 1/b1=0.02% [7], Himantormia lugubris 1/b1=1.11%, and for Caloplaca regalis 1/b1=1.93% [5].

4.31H NMR spectra

Fig. 6 shows 1H NMR spectra for C. aculeata thalli recorded at 300 MHz at two different hydration levels. For C. aculeata at the lower hydration level (Δm/m0=0.112) the spectrum was the superposition of the broad component coming from the solid matrix of the lichen thallus, which may be successfully fitted by Gaussian function, and two narrow components coming from water bound in thallus, both fitted by Lorentzian functions (Eq. (17)):(17)A(ν)=ASΔνGln4π/2exp[2ln4((ννG)ΔνG)2]+2AL1π[ΔνL14(ννL1)2+ΔνL12]++2AL2π[ΔνL24(ννL2)2+ΔνL22],where ΔνG, ΔνL1, and ΔνL2 are the half-widths of the NMR line,; νG, νL1, and νL2 are peak positions,; and finally AS, AL1, and AL2 are the amplitudes of the Gaussian and two Lorentzian peaks, respectively.

Fig. 6

1H NMR spectra for Cetraria aculeata (Schreb.) Fr. thalli recorded at 300 MHz; the pulse length π/2=1.5 μs. The relative mass increase was (a) ∆m/m0=0.112, and (b) ∆m/m0=0.218.

Fig. 6.

With the increased hydration level (at ∆m/m0=0.218) the tightly bound water fraction is no longer fitted, and the spectrum is fitted by the superposition of one Gaussian component, and only one Lorentzian component coming from loosely bound water fraction, according to:(18)A(ν)=ASΔνGln4π/2exp[2ln4(ννGΔνG)2]+2ALπ[ΔνL24(ννL2)2+ΔνL22],

The halfwidths of the solid line component in the Gaussian approximation decrease from ΔνG=(46.4±0.2) kHz down to (43.4±0.7) kHz, with the hydration level increased from Δm/m0=0.112 up to 0.218. Tightly bound water (broader) line component is detected only for the sample at the lower hydration level, with ΔνL1=2.238 kHz, whereas for the higher hydration level it is not detected by a fitting procedure. For the loosely bound water liquid component, the halfwidth is equal ca.ΔνL2=1.422 kHz, which is close to the value γ·ΔB0 of the linewidths broadened by the B0 inhomogeneities. Table 2 sets up the parameters of the recorded spectra.

Table 2

The parameters fitted to the 1H NMR spectra for Cetraria aculeata thalli recorded at room temperature for two different hydration levels.

Table 2.
Δm/m0νG[Hz]ΔνG[kHz]νL1Z [Hz]ΔνL1Z [kHz]νL2Z [Hz]ΔνL2Z [kHz]L/S*L/S (FID)
0.1122450±8046.4±0.21535±232.24±0⊡2821±71.27±0.020.4630.338
0.2183680±25043.4±0.71151±21.57±0.011.3480.719

4.4Proton free induction decays

Free induction decay for protons of the Cetraria aculeata thallus may be sufficiently well fitted by the superposition of one Gaussian component, with the amplitude S, coming from the solid matrix of the thallus; and one with the amplitude equal to L1 or for higher hydration levels two, L1 and L2, exponential components coming from water tightly and loosely bound on the surfaces of the thallus, respectively [3], [6], [7]:(19)FID(t)=Sexp((tT2S*)2)+L1exp(tT2L1*)+L2exp(tTSL2*)where T2S* is the proton relaxation time of the solid component taken for 1/e-value of the Gaussian solid signal, and T2L1* and T2L2* are the relaxation times of proton liquid fractions L1 and L2 respectively (Fig. 7).

Fig. 7

The proton FID function recorded at 30 MHz for Cetraria aculeata (Schreb.) Fr. thalli hydrated to the relative mass increase ∆m/m0=0.109; the superposition of Gaussian and two exponents (Eq. (14)) is fitted (dotted line); if the superposition of Abragam function ((16a), (16b), (16c)) and two exponentials (solid line)fits the characteristic “beat” pattern (better seen on the inset).

Fig. 7.

However, a more precise analysis reveals the presence of a “beat pattern” [38], [39] and in such a case the solid component of FID signal is better approximated by the Abragam function [40]:(20)S(t)=exp[(tT2S*)2]sinatat.

To estimate the error made by the Gaussian approximation, the moment expansion of the solid component of the FID signal was performed:(21)S(t)=S(1M22!t2+M44!t4M66!t6+),where Mi is the i-th moment of the absorption line, and t is the time starting from the final slope of /2 pulse.

For Abragam function the moment expansion equals [40]:(22)S(t)=1(2(T2S*)2+13a2)t22!+(12(T2S*)4+4a2(T2S*)2+15a4)t44!(120(T2S*)6+60a2(T2S*)4+6a4(T2S*)2+a67)t66!+where for second, fourth, and for sixth moment we get, respectively:(23a)M2=2(T2S*)2+13a2and:(23b)M4=12(T2S*)4+4a2(T2S*)2+15a2and:(23c)M6=120(T2S*)6+60a2(T2S*)4+6a4(T2S*)2+a67.

Hence, for the Gaussian approximation of the solid signal (a→0) we get:(24a)M4M22=3,and(24b)M6M23=15.

For the Cetraria aculeata thalli the moment expansion of the solid component of FID signal gives M2=3.9·109 s−2, M4=4.0·1019 s−4, and M6=2.5·1029 s−6, with the magnitude S=106.9. The solid line was nearly Gaussian (with M4M22=2.46), thus for the next routine measurements the solid line was approximated by a Gaussian function (Eq. (6)). In fact, the solid signal for lichen thalli is often nearly Gaussian in form [6], [7], [8].

Relaxation time for the solid component, T2S*≈21 μs, is close to the value obtained for other lichens, and different solid biological systems as lyophilized photosynthetic membranes [1], [2], lyophilized model membranes from DGDG [23], or lyophilized DNA [41], which suggests similar distribution of local magnetic fields for limited number of chemical groups forming the biological system at a submolecular scale [42].

The observed water fractions are differentiated by their mobility and thus by their binding and/or proximity to the solid thallus surfaces, which means that intracellular water as well as extracellular water fractions usually contributes to both these water fractions. The T2L1*≈120 μs of the L1 component is characteristic for tightly bound water of lichen thalli as well as of many other biological systems [e.g. 5, and the references therein]. The L2 signal with T2L2*≈560 μs, shortened by B0 inhomogeneities, comes from water loosely bound on the thallus surface and for the higher hydration level from free water fraction (Fig. 7).

The spin-spin relaxation times measured in FID experiment are shortened by B0 inhomogeneities according to [43]:(25)1T2*=1T2+γΔB02where T2 is the spin-spin relaxation time, γ is the gyromagnetic ratio, and ΔB0 is a change of magnetic field B0 within the sample. The solid and short exponential components are detected and not changed (as compared to CPMG echo train), but L2 may be an average of several loosely bound water pools, eg. extra- and intramolecular loosely bound water fraction.

4.5The hydration dependence of NMR signal

Fig. 8 shows the a parameter obtained from the Abragam function (Eq. (2)0) fitted to FIDs recorded for Cetraria aculeata thalli as a function of the hydration level. The parameter decreases slightly with the hydration level, which remains in correspondence with the change of the solid component halfwidth (a decrease from 32 kHz down to 25 kHz with hydration level increase from ∆m/m0=0.04 up to 0.2). For the higher hydration level the ‘beat pattern’ characteristic for Abragam function is no longer detected.

Fig. 8

The parameter a for Abragam function fitted to solid component signal (thallus solid matrix line widths) calculated for Cetraria aculeata (Schreb.) Fr. thalli as a function of hydration level recorded as relative mass increase expressed in units of dry mass Δm/m0.

Fig. 8.

For the hydration level below Δm/m0=0.18 a tightly bound water, L1, and a loosely bound water component, L2, are detected (Eq. (19)). For the tightly bound water component, the relaxation time T2L1*≈100 μs, whereas for the loosely bound water component the effective relaxation time equals T2L2*≈100 μs (see Fig. 9).

Fig. 9

The hydration dependence of proton FID relaxation times for Cetraria aculeata (Schreb.) Fr. Solid Gaussian, S, component – closed circles, tightly bound water, L1, component – open triangles, and loosely bound water, L2, fraction – open squares.

Fig. 9.

For Cetraria aculeata the relaxation time T2S* of the solid component (in a Gaussian approximation) does not change much with the thallus hydration level (Fig. 9), which suggests that the structure and molecular dynamics of the thallus solid matrix are not much modified by the hydration process. Thus, we used a solid signal amplitude, S, as an unit to scale the amplitudes of liquid components.

Fig. 10 shows the hydration dependence of the tightly and of loosely bound water signal amplitude in units of solid amplitude, L1/S and L2/S, respectively. The hydration level is expressed as the relative mass increase in units of dry mass of the thallus, Δm/m0. The tightly bound water signal, L1/S, does not change within the measured hydration range, which suggests that it has already saturated for a lower hydration level. The hydration dependence of mobile (loosely bound) water signal, L2/S, increases with the relative mass increase and is well fitted by a rational function.

Fig. 10

Tightly bound water component, L1/S, and loosely bound water component, L2/S, hydration dependence for Cetraria aculeata (Schreb.) Fr. The solid line is a rational function fitted (See text).

Fig. 10.

That the rational function successfully models the dependence, suggests the application of Eq. (6) for L1/L hydration dependence, and Eq. (11) for the L/S hydration dependence for the Cetraria aculeata thallus. The combined fit of Eq. (6) and Eq. (11) performed for γ=0.63±0.08 is presented on Fig. 11. The lowest hydration level at which L2 signal component is detected (and thus L1/L becomes less than 1) is fitted as ΔM1/m0=0.075±0.006 (see Fig. 10 for comparison). The saturation concentration of the water-soluble solid fraction is cp=0.57±0.12.

Fig. 11

The tightly bound water in units of total liquid signal, L1/L, (open triangles) and the total liquid signal in units of solid, L/S, (open circles) hydration dependence for Cetraria aculeata (Schreb.) Fr. The solid lines were calculated from Eq. (6) and from Eq. (11), respectively.

Fig. 11.

5Discussion

Although the model proposed by the Authors yields the averaged saturation concentration for water-soluble solid fraction as equal to cs=(57.3±12.0)%, which is close to the value for averaged carbohydrate and polyol saturation concentration calculated for Cetraria aculeata thallus (65.7±8.3)%, the actual concentration of carbohydrates and polyols in the dry thallus is much too small to be responsible for the nonlinearity in L/S hydration dependence.

Thus, it can be suggested than other cellular compounds are responsible for C. aculeata thallus hydration. Lichen thalli are known to contain considerable amounts of water-soluble glucans: lichenin, a hot-water-soluble linear (1→3) (1→4)-β-glucan and isolichenin, a cold-water-soluble (1→3) (1→4)-α-glucan, both with different proportions of linkage groups in diverse taxa of lichens. Honegger and Haisch [44] revealed that in fully hydrated and desiccated thalli of C. islandica, both the cortical extracellular matrix and the outer wall layer of medullary hyphae shrink dramatically during drought stress, indicating that these are major sites of water storage. Moreover authors suggested that a mycobiont derived, hydrophilic wall surface layer prevents the thallus interior from becoming waterlogged at full hydration. Similar properties, such as solubility and rheological behaviour in the solution and gel states, indicate cereal β-glucans, occurring in grasses (among them also cereals), lower plants and fungi [45], [46]. In seeds glucans play a role of storage material, which is degraded during seed germination [47]. Thus, it can be suggested that lichenin and isolichenin are also hydrolyzed in hydrated lichen thalli, according to the metabolic/physiological demand of cells. However, the content of glucans in C. aculeata was not analyzed in our study. Therefore we can only speculated that during thallus rehydration, causing the return of lichenized fungus from the cryptobiotic state, the additional pool of monosaccharides is supplied by the enzymatic action of endoglucanases. In such situation, our results could be a confirmation of the effect of the enzymatic action on lichen polysaccharides resulted in increase in monosaccharide concentration in lichen thallus returning from cryptobyosis state during rehydration. That could be a reason why L/S hydration dependence is well fitted by our model. Moreover, the increase in the levels of monosaccharides (releasing from glucans) and polyols (as a product of following algal and fungal cells metabolism) can support osmotic potential of cells leading to further water uptake [48].

In horse chestnut Aesculus hippocastantum most of water-soluble solid fraction consists mainly of sucrose. The dissolution process occurs without enzymatic action. The hydration level at which the total portion of the water-soluble solid fraction totally dissolves is directly observed in horse chestnut bast. It equals ΔM/m0=0.180, which responds for the total amount of water-soluble solid fraction equal to ΔMc/m0=0.325. Also the NMR signal from tightly bound water fraction, L1, is detected in wide range of hydration [3].

Thus, we recalculated the data for horse chestnut bast and fitted Eq. (6) to the hydration dependency of L1/L, and Eq. (11) to the hydration dependence of L/S (see Fig. 12(a)). The combined fit of these two Equations yields cs=(74.5±5.1)%, which is a little higher than 64.3% obtained using more precise model applied by them [3], and higher than those for sucrose (cs=68%). For this fitting we used γ=0.59, which is the averaged value for biological carbohydrates, as the polyol fraction for horse chestnut bast is negligible. The difference in fitted values between our and their model is a measure of real accuracy of our model. However, it may be applied in case the limit of total water-soluble solid fraction dissolution is beyond the measured hydration level range or if it does not exist at all as it is in case of germinating seeds.

Fig. 12

a. The tightly bound water in units of total liquid signal, L1/L, (open triangles) and the total liquid signal in units of solid, L/S, (open circles) hydration dependence for horse chestnut (Aesculus hippocastanum L.) bast. The solid lines were calculated from Eq. (6) and from Eq. (11), respectively. b. The tightly bound water in units of total liquid signal, L1/L, (open triangles) and the total liquid signal in units of solid, L/S, (open circles) hydration dependence for Usnea antarctica. The solid lines were calculated from Eq. (6) and from Eq. (11), respectively.

Fig. 12

In addition, we recalculated (for γ=0.63±0.08) some already published data for other species of fruticose lichenized fungi, eg. for Usnea antarctica[6]. The combined fit to Eq. (6) and to Eq. (11), respectively, yields cs=0.81±0.04, which is close to the value obtained by us for Cetraria aculeata (Fig. 12b). This suggests a similar way of recovery of the lichen metabolism from cryptobiotic form at dehydration, i.e. by the solid carbohydrate and polyol dissolution.

6Conclusions

The hydration dependency of high power 1H NMR spectra for a dehydrated thallus of Cetraria aculeata shows two fractions of residual bound water which differ in mobility, i.e. tightly and loosely bound water, and the presence of the water soluble solid fraction in the thallus. The proposed simple model allows the quantitative description of that effect and yields the saturation concentration of the discovered water soluble solid fraction. The applicability of proposed model may also extend to other tissues/organisms at drastically low hydration level, as e.g. horse chestnut bast, other species of lichenized fungi or germinating seeds.

Footnotes

Appendix ATransparency document associated with this article can be found in the online version at doi:10.1016/j.bbrep.2016.04.010.

Acknowledgements

The research was carried out with the equipment purchased thanks to the financial support of theEuropean Regional Development Fundin the framework of the Polish Innovation Economy Operational Program (contract no. POIG.02.01.00-12-023/08) and also financed by the Polish Ministry of Science and Higher Education (MNiSW, contract no. 7150/E-338/M/2015).

Appendix ATransparency document

Supplementary material
.

References

  • 1. HarańczykH.LejaA.StrzałkaK.The effect of water accessible paramagnetic ions on subcellular structures formed in developing wheat photosynthetic membranes as observed by NMR and by sorption isothermActa Phys. Pol.A1092006389398[Google Scholar]
  • 2. HarańczykH.LejaA.Jemioła-RzemińskaM.StrzałkaK.Maturation processes of photosynthetic membranes observed by proton magnetic relaxation and sorption isothermActa Phys. Pol.A1152009526532[Google Scholar]
  • 3. HarańczykH.WęglarzW.P.SojkaZ.The investigation of hydration processes in horse chestnut (Aesculus hippocastanum, L.) and pine (Pinus silvestris, L.) bark and bast using proton magnetic relaxationHolzforschung531999299310[Google Scholar]
  • 4. HarańczykH.StrzałkaK.JasińskiG.Mosna-BojarskaK.The initial stages of wheat (Triticum aestivum, L.) seed imbibition as observed by proton nuclear magnetic relaxationColloids Surf.A11519964754[Google Scholar]
  • 5. HarańczykH.On water in extremely dry biological systems2003WUJKraków
  • 6. HarańczykH.PaterŁ.NowakP.BaciorM.OlechM.A.Initial phases of Antarctic Ramalina terebrata Hook f. & Taylor thalli rehydration observed by proton relaxometryActa Phys. Pol. A12120124784822012[Google Scholar]
  • 7. HarańczykH.BaciorM.OlechM.A.Deep dehydration of Umbilicaria aprina thalli observed by proton NMR and sorption isothermAntarct. Sci.202008527535[Google Scholar]
  • 8. HarańczykH.BaciorM.JastrzębskaP.OlechM.A.Deep dehydration of Antarctic lichen Leptogium puberulum Hue observed by NMR and sorption isothermActa Phys. Pol.A1152009516520[Google Scholar]
  • 9. HarańczykH.NowakP.BaciorM.LisowskaM.MarzecM.FlorekM.OlechM.A.Bound water freezing in Umbilicaria aprina from continental AntarcticaAntarct. Sci.242012342352[Google Scholar]
  • 10. BartákM.VácziP.HájekJosefSmyklaJ.Low-temperature limitation of primary photosynthetic processes in Antarctic lichens Umbilicaria antarctica and Xanthoria elegansPolar Biol.3120074751[Google Scholar]
  • 11. HintonH.E.A new chironomid from Africa, the larva of which can be dehydrated without injuryProc. Zool. Soc. Lond.1211951371380[Google Scholar]
  • 12. HintonH.E.Cryptobiosis in the larva of Polypedilum vanderplanki Hint. (Chironomidae)J. Insect Physiol.51960286300[Google Scholar]
  • 13. WatanabeM.KikawadaT.MinagawaN.YukuhiroF.OkudaT.Mechanism allowing an insect to survive complete dehydration and extreme temperaturesJ. Exp. Biol.205200227992802[PubMed][Google Scholar]
  • 14. WatanabeM.KikawadaT.OkudaT.Increase of internal ion concentration triggers trehalose synthesis associated with cryptobiosis in larvae of Polypedilum vanderplankiJ. Exp. Biol.206200322812286[PubMed][Google Scholar]
  • 15. SakuraiM.FurukiT.AkaoKen.-ichi.TanakaD.NakaharaY.KikawadaT.Vitrification is essential for anhydrobiosis in an African chironomid, Polypedilum vanderplankiProc. Natl. Acad. Sci. USA105200850935098[PubMed][Google Scholar]
  • 16. ShimizuT.KanamoriY.FurukiT.KikawadaT.OkudaT.TakahashiT.MiharaH.SakuraiM.Dessiccation-induced structuralization and glass formation of group 3 late embryogenesis abundant protein model peptidesBiochemistry49201010931104[PubMed][Google Scholar]
  • 17. HegartyT.W.The physiology of seed hydration and dehydration, and the relation between water stress and the control of germination: a reviewPlant Cell Environ.11978101119[Google Scholar]
  • 18. LushW.M.GrovesR.H.Germination, emergence and surface establishment of wheat and ryegrass in response to natural and artificial hydration-dehydration cyclesAust. J. Agric. Res.321981731739[Google Scholar]
  • 19. BewleyJ.D.BlackM.Physiology and Biochemistry of Seeds in Relation to Germination1978Springer-VerlagBerlin, Heidelberg
  • 20. HamadaN.OkazakiK.ShinozakiM.Accumulation of monosaccharides in lichen mycobionts cultured under osmotic conditionsBryologist971994176179[Google Scholar]
  • 21. MelickD.R.SeppeltR.D.The effect of hydration on carbohydrate levels, pigment content and freezing point of Umbilicaria decussata at a continental Antarctic localityCryptogam. Bot.41994212217[Google Scholar]
  • 22. HájekJ.VácziP.BartákM.Photosynthetic electron transport at low temperatures in the green algal foliose lichens Lasalia pustulata and Umbilicaria hirsuta affected by manipulated levels of ribitolPhotosynthetica472009199205[Google Scholar]
  • 23. HarańczykH.BaciorM.JamrózJ.Jemioła-RzemińskaM.StrzałkaK.Rehydration of DGDG (digalactosyl diacylglicerol) model membrane lyophilizates observed by NMR and sorption isothermActa Phys. Pol.A1152009521525[Google Scholar]
  • 24. WitekM.WęglarzW.P.de JongL.van DalenG.BlonkJ.C.G.HeussenP.Van VelzenE.Van AsH.van DuynhovenJ.The structural and hydration properties of heat-treated rice studied at multiple length scalesFood Chem.120201010311040[Google Scholar]
  • 25. KhannaR.RajanS.GovindjeeGutowskyH.S.Effects of physical and chemical treatments on chloroplast manganese. NMR and ESR studiesBiochim. Biophys. Acta72519831018[Google Scholar]
  • 26. MurthyN.S.WorthingtonC.R.X-ray diffraction evidence for the presence of discrete water layers on the surface of membranesBiochim. Biophys. Acta10621991172176[PubMed][Google Scholar]
  • 27. de BaatW.C.Sur l’acide dithionique et ses selsRecl. Trav. Chim. Pays-Bas451926237244[Google Scholar]
  • 28. MeylanW.M.HowardP.H.BoethlingR.S.Improved method for estimating water solubility from octanol/water partition coefficientEnviron. Toxicol. Chem.151996100106[Google Scholar]
  • 30. MullinJ.W.Crystallisation1972ButterworthLondon425
  • 32. HigashiyamaT.Novel functions and applications of trehalosePure Appl. Chem.74200212631269[Google Scholar]
  • 33. LahutaL.B.ŚwięcickiW.DzikT.GóreckiR.J.HorbowiczM.Feeding stem-leaf-pod explants of pea (Pisum sativum L.) with d-chiro-inositol or d-pinitol modifies composition of α-d-galactosides in developing seedsSeed Sci. Res.202010213221[Google Scholar]
  • 34. GaffD.F.Desiccation tolerant vascular plants of Southern AfricaOecologia31197795109[PubMed][Google Scholar]
  • 35. WęglarzW.P.HarańczykH.Two-dimensional analysis of the nuclear relaxation function in the time domain: the program CracSpinJ. Phys. D Phys33200019091920[Google Scholar]
  • 36. BrunauerS.EmmettP.H.TellerE.Adsorption of gases in multimolecular layersJ. Am. Chem. Soc.601938309319[Google Scholar]
  • 37. DentR.W.A multilayer theory for gas sorption. Part I: Sorption of a single gasText. Res. J.471977145152[Google Scholar]
  • 38. WeglarzW.P.PeemoellerH.RudinA.Characterization of annealed isotactic polypropylene in the solid state by 2D time-domain 1H NMRJ. Polym. Sci. B: Polym. Phys.38200024872506[Google Scholar]
  • 39. DerbyshireW.Van Den BoschM.Van DusschotenD.MacNaughtanW.FarhatI.A.HemmingaM.A.MitchellJ.R.Fitting of the beat pattern observed in NMR free-induction decay signals of concentrated carbohydrate-water solutionsJ. Magn. Res1682004278283[Google Scholar]
  • 40. AbragamA.A.The Principles of Nuclear Magnetism1961Clarendon Press,Oxford
  • 41. HarańczykH.CzakJ.NowakP.NiziołJ.Initial phases of DNA rehydration by NMR and sorption isothermActa Phys. Pol.A1172010257262[Google Scholar]
  • 42. PintarM.M.Some considerations of the round table subjectMagn. Reson. Imaging91991753754[Google Scholar]
  • 43. TimurA.Pulsed nuclear magnetic resonance studies of porosity, movable fluid, and permeability of sandstonesJ. Pet. Technol.211969775786[Google Scholar]
  • 44. HoneggerR.HaischA.Immunocytochemical location of the (1→3) (1→4)-β-glucan lichenin in the lichen-forming ascomycete Cetraria islandica (Icelandic moss)New Phytol.1502001739746[Google Scholar]
  • 45. LazaridouA.BiliaderisC.G.Molecular aspects of cereal β-glucan functionality: Physical properties, technological applications and physiological effectsJ. Cereal Sci.462007101118[Google Scholar]
  • 46. BurtonR.A.FincherG.B.(1,3;1,4)-β-d-Glucans in cell walls of the Poaceae, lower plants, and fungi: a tale of two linkagesMol. Plant22009873882[PubMed][Google Scholar]
  • 47. CorderA.M.HenryR.J.Carbohydrate-degrading enzymes in germinating wheatCereal Chem.661989435439[Google Scholar]
  • 48. GustavsL.EggertA.MichalikD.KarstenU.Physiological and biochemical responses of green microalgae from different habitats to osmotic and matric stressProtoplasma2432010314[PubMed][Google Scholar]
Collaboration tool especially designed for Life Science professionals.Drag-and-drop any entity to your messages.