Sie sind auf Seite 1von 7

Available online at www.sciencedirect.

com

Spectrochimica Acta Part A 70 (2008) 270–276

An ATR-FTIR study of different phosphonic acids in aqueous solution


Marı́a C. Zenobi ∗ , Carina V. Luengo, Marcelo J. Avena, Elsa H. Rueda
Departamento de Quı́mica, Universidad Nacional del Sur, Avda. Alem 1253, (B8000CPB) Bahı́a Blanca, Argentina
Received 15 February 2007; received in revised form 20 July 2007; accepted 27 July 2007

Abstract
An ATR-FIR study of the vibrational spectra of 1-hydroxyethane-1,1 -diphosphonic acid (HEDP), nitrilotris(methylenephosphonic acid) (NTMP)
and N,N-bis(2-hydroxyethyl)aminomethylphosphonic acid (BHAMP) in aqueous solution is presented. The study was performed in the range of
pH from 5 to 9, and bands assignments are given in the 2000–890 cm−1 range. However, as phosphonates display bands due to the P O stretching
vibration mainly in the 900–1200 cm−1 range, the study is focused in this midinfrared region, which shows important changes as the pH changes,
specially the ν(P OH) at ∼925 cm−1 and ν(PO3 2− ) at ∼970 cm−1 vibrations. IR analyses give also evidences for the zwitterionic nature of BHAMP
and NTMP in solution with a strong indication that the zwitterion in both compounds remains intact throughout the pH range investigated. The
successive protonation steps with the decrease of pH were evidenced in the IR spectra of the three studied phosphonates.
© 2007 Elsevier B.V. All rights reserved.

Keywords: ATR-FTIR; Phosphonic acid; BHAMP; HEDP; NTMP; Band assignment

1. Introduction thoroughly reviewed [5–10]. Much less is known about the


fate and behaviour of the corresponding phosphonates in the
Phosphonic acids and their derivatives are of interest due environment. In recent reviews, discussions about environmen-
to their structural variety and great economic importance. Their tal chemistry of phosphonates, and their toxicology and risk
high water solubility, chemical stability and effective capacity to assessment were provided by several authors [4,11,12]. It is well
disperse or flocculate solids, as well as their peroxy stabilisation known that phosphonates posses a very high ability to form com-
properties promote their wide use, e.g., in water cooling systems, plexes with transition metals [13] and they also have a strong
in bleaching baths or as scale inhibitors in deflocculation agents tendency to adsorb onto a variety of surfaces [14–16]. To fully
[1]. Biphosphonates are also used in medicine to treat various understand the chemical processes that occur in groundwater,
bone and calcium metabolism diseases [2,3]. soils and surface waters, it is important to examine the role of
In some instances phosphonates are discharged into wastewa- the solid–water interface as well as interactions in the aqueous
ter treatment plants, while in others they are discharged directly phase. At the present there is some information in the literature
into natural waters. Due to the presence of natural phospho- regarding infrared spectroscopy of glyphosate, amino methyl
nates in the environment, bacteria have evolved the ability to phosphonic acid, methyl phosphonic acid and phenylphospho-
metabolise phosphonates and to use them as a source of nutri- nate [17–22]. However, no information can be found regarding
ents. In fact, those bacteria able of cleaving the C P bond can use di- or polyphosphonic acids. A detailed knowledge of the
phosphonates as a phosphorus source for growth. However, the vibrational spectra of these substances in aqueous solution is
possibility of other more accessible phosphorus sources on the necessary before attempting to determine their behaviour in
phosphonates uptake and degradation are of great environmental further adsorption studies.
importance [4]. In this work, an in situ ATR-FTIR spectroscopy study
Environmental concentrations, usage, biodegradation and of three phosphonic acids in aqueous solution at differ-
toxicology of carboxylates and aminocarboxylates have been ent pH is presented. Two of the phosphonic acids studied
here, 1-hydroxyethane-1,1 -diphosphonic acid (HEDP, LH4 )
and nitrilotris(methylenephosphonic acid) (NTMP, LH6 ), are
∗ Corresponding author. Tel.: +54 291 4595101x3579; fax: +54 291 4595160. common components of detergents and can be found in
E-mail address: mzenobi@criba.edu.ar (M.C. Zenobi). the environment [23,24]. The other compound, N,N-bis(2-

1386-1425/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.saa.2007.07.043
M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276 271

hydroxyethyl)aminomethylphosphonic acid (BHAMP, LH2 ),


was chosen for comparison because it is a monophosphonic acid
with a hydroxyl group and an amine group similar as HEDP and
NTMP, respectively.

2. Materials and methods

2.1. Materials

All solutions were prepared using double-distilled deionized


water and reagent-grade chemicals. BHAMP was provided by
A.R. Mass Chemical, California (USA). Gador laboratories pro-
vided HEDP, which was purified by removing free phosphate
ions with the aid of an anion-exchange resin (Amberlite IRA-
400). NTMP in the acid form was obtained from Aldrich. The
chemical structure of the three studied compounds is shown in
Fig. 1.

2.2. ATR-FTIR spectroscopy

The ATR-FTIR spectra were obtained with 4 cm−1 reso-


lution on a Nicolet Magna 560 spectrometer equipped with
DTGS detector using a ZnSe crystal (45◦ angle of incidence)
as the internal reflection element. Each spectrum is the result
of 256 co-added interferograms, in the range 4000–800 cm−1 .
The working temperature in these experiments was 25 ◦ C. The Fig. 2. ATR-FTIR spectra of 1 × 10−2 M phosphonic acids solutions at pH 9.
FTIR spectra of phosphonic acid solutions were recorded after
filling the ATR cell with 1 × 10−2 M solutions in the pH range
5–9. Double-distilled deionized water was used for background Fig. 2 shows the spectra of the studied phosphonates
spectra. at pH 9 where all phosphonates display a high deprotona-
tion level. In the 2000–1200 cm−1 range appears one main
3. Results and discussion band at 1351–1359 cm−1 corresponding to vibration δOH of
the alcoholic group present in BHAMP and HEDP spectra.
3.1. ATR-FTIR spectra The absence of this band on the NTMP spectrum confirms
the above-mentioned assignation. Small bands observed at
Although the ATR spectra were recorded in the 4000– 1600–1500 cm−1 range are due to vibrations of dissolved car-
800 cm−1 range, only bands in the 2000–900 cm−1 were ana- bonate and water [25,26]. The bands on the 1200–900 cm−1
lyzed, because this is the range where the main bands of region are those related with the change in the protonation of
functional groups of BHAMP, HEDP and NTMP appear. the phosphonic group, which are studied as a function of pH in
the following sections [17,19–21].

3.2. BHAMP

Variations in the pH of the solutions induce protonation or


deprotonation of the oxoacid ligand thereby resulting in a change
in the ligand symmetry. This change affects the position and the
number of bands in the infrared spectra [27,28]. Consequently,
the peak assignment for the infrared spectra of the BHAMP
solution can be mainly based on the species distribution in
BHAMP solutions at different pH values. Based on the dissoci-
ation constant of BHAMP (Table 1) the distribution of species
as a function of pH can be determined (Fig. 3).
ATR-FTIR spectra of the BHAMP in aqueous solution as a
function of pH are illustrated in Fig. 4. To obtain more detailed
information from this set of spectra, we used the program Peak
Fig. 1. Structures of the studied phosphonic acid: BHAMP, HEDP and NTMP. Fit 4.0 to deconvolute them into several peaks.
272 M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276

Table 1
Acidity constants of phosphonic acids (I = 0.1 mol L−1 (KNO3 ), 25 ± 0.5 ◦ C)
pKa1 pKa2 pKa3 pKa4 pKa5 pKa6

BHAMPa 4.98 9.46


HEDPb 1.43 2.70 7.02 11.20
NTMPc 0.50 1.60 4.59 5.90 7.22 12.50
a Taken from Ref. [29].
b Taken from Ref. [30].
c Taken from Ref. [13].

The ATR-FTIR spectrum of BHAMP at pH 9 corresponds


to a mix of the protonated and unprotonated (LH− and L2− )
ions, which are, respectively, the (R)2 NH+ CH2 (PO3 )2− and
(R)2 N CH2 (PO3 )2− species. At this pH two bands at 1097
and 978 cm−1 dominate the spectrum, what could be assigned to
νas (PO3 2− ) and νs (PO3 2− ) vibrations, respectively. The decon-
volution of these spectra shows another band at 1064 cm−1
attributed to asymmetric vibration of fully deprotonated P O.
These assignments were carried out on the basis of the studies
performed by Barja and Dos Santos Afonso [18] and Sheals et
al. [20], who studied glyphosate solutions in water. They found
that the band at 1095 cm−1 , which was attributed to a zwit-
terionic structure involving intramolecular hydrogen bonding
between P O and NH2 + in LH− , shifted by around 30 cm−1
towards lower wavenumbers (1067 cm−1 ) upon deprotonation
of the amino group to produce the L2− species of glyphosate.
This shift in the band position appears to be a consequence of
the increase in the total negative charge of the glyphosate unit
[18,20].
At pH 8 only the (R)2 NH+ CH2 (PO3 )2− species (LH− )
of BHAMP is present at significant concentrations. There-
fore, the band corresponding to L2− is not present in the
spectrum and only the two bands at 1098 and 980 cm−1
appear. Very similar spectra are found at pH 7 and 6, where
(R)2 NH+ CH2 (PO3 )2− is the predominant species.
At pH 5 two new bands at 1192 and 924 cm−1 appear
in the spectrum. The band at 1192 cm−1 corresponds to the
asymmetric stretching vibration of P O (νas HPO3 − ). Accord-

Fig. 4. Deconvoluted ATR-FTIR spectra of 1 × 10−2 M BHAMP in aqueous


solution at different pH values: (circle) experimental spectra, (solid line) fitted
spectra and (dotted line) individual peaks.

ing to the literature [18,20] a band should also appear at


∼1082 cm−1 corresponding to symmetric stretching vibration
of P O (νs HPO3 − ). By deconvolution, we found that this band is
somewhat masked by the band of νas (PO3 2− ). On the other hand,
the band at 924 cm−1 corresponds to the stretching vibration of
P OH. These assignments at 1192 and 924 cm−1 are in agree-
Fig. 3. Distribution diagram for BHAMP at 25 ◦ C. LH2 represents
(R)2 NH+ CH2 (HPO3)− , whereas LH− and L2− represent the species gen- ment with previous assignments for phosphates [27,31–33].
erated after the successive deprotonation of phosphonic and amino groups, The peak assignments for the spectra of the aqueous BHAMP
respectively. solutions are summarized in Table 2.
M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276 273

Table 2
Tentative assignments of ATR-FTIR peaks for BHAMP in aqueous solution
Wavenumber (cm−1 ) Assignment

1192 νas P O of HPO3 −


1091–1098 νas P O of PO3 2−
1082 νs P O of HPO3 −
1064 νas P O of PO3 2−
978–980 νs P O of PO3 2−
924 νas P OH of HPO3 −

3.3. HEDP

The distribution of species of HEDP at different pH are shown


in Fig. 5, as calculated from the proton dissociation constants
listed in Table 1. ATR-FTIR spectra of the HEDP solutions as a
function of pH are illustrated in Fig. 6. As it can be observed, the
spectra of HEDP are somewhat more complicated than those of
BHAMP due to the presence of two phosphonic groups instead
of one.
At pH 9 the only species of HEDP at significant concen-
trations is LH3− . The ATR-FTIR spectrum at this pH shows
several bands: three bands at 1099, 1077 and 968 cm−1 that can
be assigned to νas (PO3 2− ), νas (PO3 2− ) and νs (PO3 2− ), respec-
tively, and bands at 1142 and 1044 cm−1 that correspond to
νas (HPO3 − ) and νs (HPO3 − ), respectively. The presence at pH
9 of two bands corresponding to νas (PO3 2− ) could be due to the
coexistence of two phosphonic groups in LH3− , where only one
of them is protonated, so that each phosphonic group exhibit
different symmetry. The appearance of the band at higher fre-
quency (1099 cm−1 ) is a consequence of the increase in the P O
bond strength of the two unprotonated P O groups as an indirect
effect of the protonation of the third P O group [21].
At pH 8, the spectrum is very similar to that at pH 9, which is
consistent with the fact that mainly LH3− is present in the solu-
tion. At pH 7 a new protonated species (LH2 2− ) appears in the
solution producing a change in the spectrum. At this pH exists
the possibility of the presence of (HPO3 )− R (HPO3 )− and
(H2 PO3 ) R (PO3 )2− , along with LH3− . Therefore, in the spec-
trum the bands corresponding to each species should be present.

Fig. 6. Deconvoluted ATR-FTIR spectra of 1 × 10−2 M HEDP in aqueous solu-


tion at different pH values: (circle) experimental spectra, (solid line) fitted spectra
and (dotted line) individual peaks.

By deconvolution, a good fit was obtained with eight bands: the


bands at 1146 and 1050 cm−1 could be assigned to νas (HPO3 − )
and νs (HPO3 − ); the band at 1174 cm−1 corresponds to ν(P O);
the bands at 1101 and 1075 cm−1 are assigned to νas (PO3 2− ) and
that at 969 cm−1 to νs (PO3 2− ), whereas the bands at 1213 and
931 cm−1 are attributed to δ(P OH) and ν(P OH), respectively.
Fig. 5. Distribution diagram for HEDP at 25 ◦ C. LH4 represents At pH 6 the same pattern of bands than at pH 7 appears, but
R,R C (PO3 H2 )2 , whereas LH3 − , LH2 2− , LH3− and L4− represent it is observed a decrease in the bands corresponding to PO3 2−
the species generated after successive deprotonation of phosphonic groups. (1111, 1085 and 970 cm−1 ) and an increase in the bands assigned
274 M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276

Table 3
Tentative assignments of ATR-FTIR peaks for HEDP in aqueous solution
Wavenumber (cm−1 ) Assignment

1213–1218 δP OH of HPO3 −
1174–1177 νP O
1142–1146 νas P O of HPO3 −
1099–1111 νas P O of PO3 2−
1075–1085 νas P O of PO3 2−
1044–1057 νs P O of HPO3 −
968–971 νs P O of PO3 2−
931–938 νas P OH of HPO3 −

to HPO3 − (1144 and 1057 cm−1 ), POH (1218 and 933 cm−1 )
and P O (1176 cm−1 ). At pH 5, the same bands appear but there
is an increase in the intensity of the bands corresponding to the
protonated species.
Gong [33] has shown that the spectra of the linear polyphos-
phate solutions vary markedly with pH. This can be explained
by the increasing protonation of the phosphate species with
decreasing pH that affect the symmetry and the infrared vibra-
tional frequency of the aqueous orthophosphate species. In that
work, Gong [33] established that as the protonation increases,
some of the νas (P O) change into νas (P OH) with much
lower frequency, whereas the νas frequency of the remain-
ing free (unbound) P O increases significantly. Similarly, with
increasing protonation, the bands of νas (P O) in the spectra of
polyphosphate solutions all decrease in intensity and shift to
higher frequency, because some of the νas (P O) have become
νas (P OH).
The peak assignments for the spectra of the aqueous HEDP
solutions are summarized in Table 3.

3.4. NTMP

The distribution of species of NTMP at different pH is shown


in Fig. 7, as calculated from the proton dissociation constants
listed in Table 1. ATR-FTIR spectra of the NTMP solutions as
a function of pH are illustrated in Fig. 8.

Fig. 8. Deconvoluted ATR-FTIR spectra of 1 × 10−2 M NTMP in aqueous solu-


tion at different pH values: (circle) experimental spectra, (solid line) fitted spectra
and (dotted line) individual peaks.

At pH 9 two bands dominate the spectrum at 1101


and 970 cm−1 corresponding to vibrations νas (PO3 2− ) and
νs (PO3 2− ). The main species at this pH is LH5− , where all the
phosphonic groups are unprotonated. The band at 1101 cm−1
is attributed to a zwitterionic structure involving intramolecu-
lar hydrogen bonding between P O and NH+ in LH5− . This
Fig. 7. Distribution diagram for NTMP at 25 ◦ C. LH6 represents
assignment agrees with results found by Sheals et al. [21]
(HPO3 )− CH2 NH+ (PO3 H2 )2 , whereas LH5 − , LH4 2− , LH3 3− LH2 4− ,
LH5− and L6− represent the successive deprotonation of phosphonic and amino for glyphosate. The absence of other bands was attributed
groups. to the delocalization of the electrons over the terminal oxy-
M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276 275

Table 4 ICET for finantial support. We thank Olga Pieroni (Universidad


Tentative assignments of ATR-FTIR peaks for NTMP in aqueous solution Nacional del Sur) for her help with the ATR measurements. C.L.
Wavenumber (cm−1 ) Assignment thanks CONICET for the doctoral fellowship.
1220–1228 δP OH of HPO3 −
1170–1180 νas P O of HPO3 − References
1101–1110 νas P O of PO3 2−
1079–1080 νas P O of PO3 2− [1] J. Svara, N. Weferling, T. Hofmann, Organic Phosphorus Compounds, in:
1050–1062 νs P O of HPO3 − Ullmann’s Encyclopedia of Industrial Chemistry, online edition, Wiley-
970–975 νs P O of PO3 2− VCH, 2002.
921–926 νas P OH of HPO3 − [2] G.A. Fromm, F. Schajowicz, C. Casco, G. Ghiringhelli, C. Mautalen, The
treatment of Paget’s bone disease with sodium etidronate, Am. J. Med. Sci.
277 (1979) 29–37.
[3] H. Fleisch, Bisphosphonates: a new class of drugs in diseases of bone and
gen atoms which makes identical all the terminal P O bonds calcium metabolism, Recent Results Cancer Res. 116 (1989) 1–28.
[34]. [4] B. Nowack, Environmental chemistry of phosphonates, Water Res. 37
When the pH decreases to 8, three bands appear at 1102, (2003) 2533–2546.
1079 and 970 cm−1 corresponding to νas (PO3 2− ), νas (PO3 2− ) [5] J.L. Means, C.A. Alexander, The environmental biogeochemistry of chelat-
ing agents and recommendation for the disposal of chelated radioactive
and νs (PO3 2− ), respectively, and two new bands at 1180 and wastes, Nucl. Chem. Waste Manag. 2 (1981) 183–196.
1054 cm−1 are present in the spectrum, that could be assigned to [6] W.L. Anderson, W.E. Bishop, R.L. Campbell, A review of the environ-
νas (HPO3 − ) and νs (HPO3 − ), respectively, due to the appearance mental and mammalian toxicology of nitrilotriacetic acid, CRC Crit. Rev.
of the protonated LH2 4− species. Similar to HEDP, the existence Toxicol. 15 (1985) 1–102.
of three phosphonic groups where only one of them is proto- [7] F. Frimmel, Complexing acids, in: M.J. Schwuger (Ed.), Detergents in the
Environment, Marcel Deckker, Inc., New York, 1997, pp. 289–312.
nated in the LH2 4− species produces P O groups of different [8] B. Nörtemann, Biodegradation of EDTA, Appl. Microbiol. Biotechnol. 51
symmetry promoting the formation of two bands corresponding (1999) 751–759.
to νas (PO3 2− ) [21]. [9] B. Nowack, Environmental chemistry of aminopolycarboxylate chelating
At pH 7 two new bands appear at 1220 and 925 cm−1 , agents, Environ. Sci. Technol. 36 (2002) 4009–4016.
attributed to δ(P OH) and ν(P OH), respectively. At pH 6 there [10] T.P. Knepper, Synthetic chelating agents and compounds exhibiting com-
plexing properties in the aquatic environment, Trends Anal. Chem. 22
is a similar concentration of both species (LH2 4− and LH3 3− , (2003) 708–724.
Fig. 7), leading to a diminution of the bands corresponding to [11] B. Nowack, Environmental chemistry of phosphonic acids, in: E. Valsami-
PO3 2− (1105, 1080 and 975 cm−1 ), and an increase of the bands Jones (Ed.), Phosphorous in Environmental Technologies: Principles and
assigned to HPO3 − (1170 and 1050 cm−1 ) as compared to the Applications, IWA Publishing, London, 2004, pp. 147–173.
spectrum at pH 7. The spectrum at pH 5 shows the same pattern [12] J. Jaworska, H. Van Genderen-Takken, A. Hanstveit, E. van de Plassche, T.
Feijtel, Environmental risk assessment of phosphonates, used in domestic
of bands than at pH 6. laundry and cleaning agents in The Netherlands, Chemosphere 47 (2002)
An increase in the intensity of the peaks corresponding 655–665.
to asymmetric stretching vibrations of P O bond and P OH [13] V. Deluchat, J.C. Bollinger, B. Serpaud, C. Caullet, Divalent cations spe-
bond in HPO3 − is observed when pH decreases from 9 to ciation with three phosphonate ligands in the pH-range of natural waters,
5. This is attributable to the increase in the amount of proto- Talanta 44 (1997) 897–907.
[14] B. Nowack, A.T. Stone, Adsorption of phosphonates onto the
nated phosphonate groups in the solution, which promote the goethite–water interface, J. Colloid Interface Sci. 214 (1999) 20–30.
electron polarization and lead to the increase in the force con- [15] M.C. Zenobi, L. Hein, E. Rueda, The effect of 1-hydroxyethane-(1,1-
stant of the P O bond [28]. This estimative assignment is in diphosphonic acid) on the adsorptive partitioninc of metal ion onto
agreement with the vast literature on IR spectroscopy of phos- ␥-AlOOH, J. Colloid Interface Sci. 284 (2005) 447–454.
phates/phosphonates [18–22,27,28,32–34]. [16] M.C. Zenobi, E.H. Rueda, Adsorption of Me(II), HEDP, and Me(II)-HEDP
onto boehmite at non stoichiometric Me(II)-HEDP concentrations, Envi-
The peak assignments for the spectra of the aqueous NTMP ron. Sci. Technol. 40 (2006) 3254–3259.
solutions are summarized in Table 4. [17] P. Persson, E. Laiti, L.O. Öhmann, Vibration spectroscopy study of phe-
In summary, from the results presented in the figures above, nilphosphonate at the water–aluminum (Hydr) oxide interface, J. Colloid
it is clear that the 900–1200 cm−1 IR region is very impor- Interface Sci. 190 (1997) 341–349.
tant for studying the protonation–deprotonation of phosphonate [18] B.C. Barja, M. Dos Santos Afonso, An ATR-FTIR study of glyphosate and
its Fe(III) complex in aqueous solution, Environ. Sci. Technol. 32 (1998)
groups of the studied compounds in aqueous solutions, spe- 3331–3335.
cially the ν(P OH) at ∼925 cm−1 and ν(PO3 2− ) at ∼970 cm−1 [19] B.C. Barja, M.I. Tejedor-Tejedor, M.A. Anderson, Complexation of
vibrations. The assigned bands are of interest as they are methylphosphonic acid with the surface of goethite particles in aqueous
expected to be diagnostic bands in the studies of the interac- solution, Langmuir 15 (1999) 2316–2321.
tions of phosphonic acid with metal ions and solid surfaces in [20] J. Sheals, P. Persson, B.I.R. Hedman, EXAFS spectroscopic studies of
glyphosate protonation and copper(II) complexes of glyphosate in aqueous
water. solution, Inorg. Chem. 40 (2001) 4302–4309.
[21] J. Sheals, S. Sjöberg, P. Persson, Adsorption of glyphosate on goethite:
molecular characterization of surface complexes, Environ. Sci. Technol.
Acknowledgments
36 (2002) 3090–3095.
[22] B.C. Barja, M. Dos Santos Afonso, Aminomethylphosphonic acid and
We are grateful to the Secretarı́a General de Ciencia y glyphosate adsorption onto goethite: a comparative study, Environ. Sci.
Tecnologı́a (SGCyT), Universidad Nacional del Sur, and CON- Technol. 39 (2005) 585–592.
276 M.C. Zenobi et al. / Spectrochimica Acta Part A 70 (2008) 270–276

[23] B. Nowack, Determination of phosphonates in natural waters by ion-pair [29] R. Karliček, J. Majer, J. Polakovičová, New complexanes. XXI. Complex-
high-performance liquid chromatography, J. Chromatogr. A 773 (1997) forming properties of the N,N-bis(2-hydroxyethyl)aminomethylpho-
139–146. sphonic acid, Chem. Zvesti 24 (1970) 161–172.
[24] B. Nowack, The behavior of phosphonates in wastewater treatment plants [30] K. Popov, H. Rönkkömäki, L.H.J. Lajunen, Critical evaluation of stability
of Switzerland, Water Res. 32 (1998) 1271–1279. constant of phosphonic acid, Pure Appl. Chem. 73 (2001) 1641–1677.
[25] R.A. Nyquist, R.O. Kagel, Infrared Spectra of Inorganic Compounds [31] A.C. Chapman, L.E. Thirlwell, Spectra of phosphorus compounds-I. The
(3800–45 cm−1 ), Academic Press, Inc., New York, 1971. infrared spectra of orthophosphates, Spectrochim. Acta 20 (1964) 937–947.
[26] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination [32] P. Persson, N. Nilsson, S. Sjöberg, Structure and bonding of orthophosphate
Compounds, Willey, New Cork, 1986. ions at the iron oxide-aqueous interface, J. Colloid Interface Sci. 177 (1996)
[27] M.I. Tejedor-Tejedor, M.A. Anderson, The protonation of phosphate on the 263–275.
surface of goethite as studied by CIR-FTIR and electrophoretic mobility, [33] W. Gong, A real-time in-situ ATR-FITIR spectroscopic study of linear
Langmuir 6 (1990) 602–611. phosphate on titania surfaces, Int. J. Miner. Process. 63 (2001) 147–165.
[28] X.-H. Guan, C. Shang, J. Zhu, G.-H. Chen, ATR-FTIR investigation on the [34] X.-H. Guan, Q. Liu, G.-H. Chen, C. Shang, Surface complexation of con-
complexation of myo-inositol hexaphosphate with aluminum hydroxide, J. densed phosphate to aluminum hydroxide: an ATR-FTIR spectroscopic
Colloid Interface Sci. 293 (2006) 296–302. investigation, J. Colloid Interface Sci. 289 (2005) 319–327.

Das könnte Ihnen auch gefallen