in copyright - non-commercial use permitted rights ...48490/eth-48490-02.pdf2-ratios of 1:10. those...

180
Research Collection Doctoral Thesis Biomass Related Salt Solutions at Hydrothermal Conditions Author(s): Reimer, Joachim Publication Date: 2015 Permanent Link: https://doi.org/10.3929/ethz-a-010578835 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: others

Post on 20-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Research Collection

Doctoral Thesis

Biomass Related Salt Solutions at Hydrothermal Conditions

Author(s): Reimer, Joachim

Publication Date: 2015

Permanent Link: https://doi.org/10.3929/ethz-a-010578835

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Page 2: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Joac

him

Rei

mer

Bio

mas

s R

elat

ed S

alt

Solu

tio

ns

at H

ydro

ther

mal

Co

nd

itio

ns

Dis

s. E

TH N

o. 2

31

04

Paul Scherrer Institut, 5232 Villigen PSI, Switzerland Tel. +41 56 310 21 11, Fax +41 56 310 21 99

www.psi.ch

Diss. ETH No. 23104

Biomass Related Salt Solutions at Hydrothermal Conditions

Joachim Reimer

Page 3: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2015

DISS. ETH NO. 23104

BIOMASS RELATED SALT SOLUTIONS AT HYDROTHERMAL CONDITIONS

A thesis submitted to attain the degree of

DOCTOR OF SCIENCES of ETH ZURICH

(Dr. sc. ETH Zurich)

presented by

JOACHIM REIMER

Dipl.-Chem., Karlsruhe Institute of Technology

born on 19.01.1987

citizen of Germany

accepted on recommendation of

Prof. Dr. Alexander Wokaun

Prof. Dr. Marco Mazzotti

Prof. Dr. Frédéric Vogel

Page 4: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise
Page 5: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

I

Für meine Familie

There is no destination,

there is only the journey!

David Gold (1980-2011)

Page 6: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

II

Page 7: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

III

Danksagung

Ich danke Prof. Dr. A. Wokaun für die Übernahme des Referats meiner Arbeit und die

anregenden Diskussionen in den Quartalsgesprächen

Mein Dank gilt weiterhin Prof. Dr. M. Mazzotti für die Übernahme des Koreferats dieser

Arbeit.

Prof. Dr. Frédéric Vogel danke ich für die Betreuung in den vergangenen Jahren und für die

Möglichkeit meine Dissertation auf einem so interessanten Themengebiet durchführen zu

können.

Ein besonderer Dank geht an Erich De Boni, den Techniker unserer Gruppe, der mir stets mit

Rat und Tat zur Seite stand und ohne den die technische Umsetzung der meisten Projekte

nicht möglich gewesen wäre. Erich, vielen Dank für die vielen Abende im „Uusgang“, den

guten Wein, die Gespräche und Diskussionen in allen Lebenslagen und für deine

Freundschaft!

Vielen Dank an das Techniker-Team, für die schnelle und unkomplizierte Unterstützung:

Marcel Hottiger, Peter Hottinger, Lukas Oberer, Lorenz Bäni.

I thank Dr. Sonia Pin for the good and bad times during our endless beamtimes at PHOENIX.

Dr. Thomas Huthwelker danke ich für die Betreuung und Unterstützung vor, während und

nach den Strahlzeiten an PHOENIX.

Many thanks also to Prof. Dr. Matthew Steele-MacInnis for teaching me MD, for the fruitful

collaboration, and for the proofreading of this thesis.

Dr. Jörg Wambach danke ich für die Möglichkeit das Raman Setup benutzen zu dürfen.

Ich danke meinen Kollegen für die Unterstützung, die gemeinsamen Mittagspausen und

Unternehmungen. Namentlich in der zeitlichen Reihenfolge des Auftretens sind dies:

Dr. Martin Schubert, Dr. Johannes Müller, Dr. Hemma Mehring geb. Zöhrer, Franziska Mayr,

Dario Wüthrich, Simon Maurer, Vera Tschedanoff, Christian Bährle, Sebastian Viereck, Frank

Pilger, Mattia Lucchini, Agnese Carino, Dominik Gschwend, David Scholz und alle anderen

Mitgliedern des LBK und der TPE Gruppe.

Ein Dank geht auch an meine Freunde, die mich während meiner Dissertation begleitet

haben: Steffen, Johanna, Jo, Franzi, Bettina, Vanessa, Heiko, Lukas, Johannes, Max, Philipp,

Patrick, Janne, Nina, Augustin, Marc und Urs.

Insbesondere danke ich Sabrina Conrad, die mich das gesamte Studium und die ersten Jahre

der Promotion begleitet hat. Vielen Dank für die gemeinsame Zeit!

Meiner Familie danke ich ganz besonders herzlich, dass sie mich auf meinem bisherigen

Lebensweg immer tatkräftig unterstützt hat!

Page 8: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

IV

Page 9: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

V

Abstract

Aqueous salt solutions at elevated pressures and temperatures play an important role in

Earth’s crust, where they contribute to heat and mass transfer and are crucial for the

formation of ore deposits, and these solutions are also used in energy production such as in

power-plants (nuclear and conventional), hydrothermal treatment of biomass, such as

supercritical water gasification (SCWG), or in the destruction of hazardous waste, like in

supercritical water oxidation process (SCWO).

At the Paul Scherrer Institute, a process is being developed to convert biomass with a high

moisture content into methane as an energy carrier, in order to contribute to meeting

mankind's future energy needs in a sustainable and CO2-neutral way.

In this process the phase behavior of the inorganic constituents of the feedstock is crucial, as

these inorganic salts may lead to plugging of the plant or poisoning/fouling of the Ru/C

catalyst. Hence, the inorganic fraction of the feedstock must be removed from the process

stream to avoid such problems. Furthermore, recovery of the salts may yield fertilizer as a

value-added product, which can be used for biomass production, closing the nutrient cycle

and making the whole process more economically feasible and attractive.

In order to optimize the process design in terms of salt separation and recovery,

fundamental questions as to the physical chemistry of aqueous salt solutions at

hydrothermal conditions must be addressed. The nature of various types of salts, their

interactions with the aqueous solvent, and the controls on phase equilibria are key facets of

these questions.

The aim of this thesis is to investigate the phase behavior and its microscopic origin of model

salt solutions, as well as the application of the gathered knowledge for the improvement

SCWG process.

Two main types of salts have been used as model compounds in these investigations. So

called type 1 salts exhibit a continuous solubility line in their binary aqueous phase diagrams,

extending from the triple point of pure water to the triple point of the pure salt. The phase

diagrams of such salts are often complicated by the appearance of a certain liquid-

Page 10: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

VI

immiscibility (then called type 1d). The second type of salts (denominated type 2) features a

discontinuous solubility line. Intersection of the gas-liquid critical line with the solubility line

leads to a region between the critical end points p and Q, where a single fluid phase exists,

typically exhibiting very low salt concentrations, in equilibrium with solid salt. In an earlier

work, the type 1(d) salts could be efficiently separated and recovered from the process

stream in the SCWG process, whereas the type 2 salts caused plugging of the plant due to

solid precipitation.

In the first part of the thesis Differential Scanning Calorimetry (DSC) was used to investigate

the phase behavior of binary and ternary model salt solutions. Using binary model salt

solutions, it was proven that DSC measurements are well suited to determine the nature of

phase transitions in hydrothermal salt solutions; namely, differentiating between

precipitation of solid salt, liquid immiscibility, and gas-liquid homogenization phenomena.

The gathered knowledge then was used to investigate the phase behavior of ternary salt

solutions, consisting of two salts and water as a solvent. In rare cases a mixture of two type 2

salts can exhibit a phase behavior of a type 1d solution, depending on the concentration and

molar ratio between the salts. This effect is possible through the stabilization of a liquid

immiscibility, which is metastable in the binary salt solutions. This was the case in the

ternary mixture NaxK2-xSO4-H2O. Mixed solutions of sodium or potassium sulfate with

magnesium sulfate did not exhibit a stabilization of the liquid immiscibility. Using carbonate

as an anion, the solutions with sodium and potassium ions only exhibited type 1(d) behavior

at Na+:K+ molar ratios lower than 10:1. Replacing sulfate by hydrogen-phosphate ions,

together with sodium or potassium as cations, led to large ranges of composition, in which

the liquid immiscibility is stabilized, up to SO42-:HPO4

2- ratios of 1:10. Those results indicate

that addition of K2HPO4 can be used to help convert otherwise type-2 aqueous salt systems

towards type 1(d) behavior, which is favorable for salt recovery. Furthermore, the phase

behavior in ternary salt solutions follows the same trends as in binary solutions; thus,

solutions containing highly charged or small-radius ions mostly exhibit phase behavior of

type 2, whereas larger, less charged ions lead to a phase behavior of type 1(d). In addition,

the number of coordination sites seems to be important in the case of polyatomic anions.

Page 11: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

VII

The second part of the thesis deals with the microscopic investigation of the salt solutions at

hydrothermal conditions. Using modulated FT-IR Raman spectroscopy on a hydrothermal

Na2SO4 solution (a type 2 solution), we found a shift in the Raman signal, which is attributed

to the free ion at ambient conditions, and a Raman signal that appears close to the

precipitation conditions, which increases in intensity upon tuning the pressure/temperature

condition closer to the solubility limit. Aided by Molecular Dynamics (MD) simulations, we

could attribute this signal to ionic clusters containing up to ~5 sulfate ions, which appear to

be the precursor of the solid phase prior to precipitation.

Replacement of a fraction of the sodium ions with potassium ions led to the disappearance

of the Raman signal of these clusters. DSC measurements revealed a strong increase in the

solubility in such solutions with increasing fraction of potassium. Further MD simulations

revealed that the exchange of sodium with potassium ions leads to a breakdown of the

clusters due to the weaker electrostatic interaction of the potassium ions with the sulfate

ions, caused by the larger ionic radius of the potassium ion compared to sodium. In addition,

the MD simulations on the ternary solutions showed chain-like polymeric structures of the

associated ions. It has previously been speculated that such chain-like structures may be

responsible for liquid immiscibility. Hence our findings support this idea, as a liquid

immiscibility is also present in the ternary NaxK2-xSO4-H2O solutions (but absent in the

relevant binaries).

In the third part of the thesis, a new design for a salt separator is presented, resulting from

considerations of the phase behavior of the salt solutions. In contrast to the design used

earlier in PSI’s SCWG process, where the feed solution is injected at the top of the vessel via

a dip-tube, we use a riser-tube to feed the biomass from the bottom into the vessel. This

leads to the formation of a pseudo-interface between a low-density fluid at the top of the

reactor and a high density fluid at the bottom. In the low-density fluid, the salts have very

low solubility, and the water/organic mixture can be withdrawn for methanation over a

catalyst. The high density fluid keeps all the inorganic salts dissolved, and a concentrated

brine can be withdrawn at the bottom of the reactor. First assessment of the new design

yielded promising results for both model solutions and real biomass.

Page 12: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

VIII

Overall it can be concluded that the phase behavior of mixtures can be successfully adjusted

by addition of suitable salt(s), which influence ion-solvent interactions in predictable ways,

based on the results of this investigation. This knowledge has led to a new salt-separator

design, which appears promising. Furthermore, the clustering tendency of ions, mediated

through their electrostatic interaction (driven mainly by charge and ionic radii), was

identified as the key for the microscopic origin of the macroscopic phase behavior.

Page 13: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

IX

Zusammenfassung

Wässrige Salzlösungen unter erhöhtem Druck und erhöhter Temperatur spielen eine

wichtige Rolle in der Erdkruste, wo sie zum Wärme- und Stofftransport beitragen und

wichtig für die Bildung von Erzlagerstätten sind. Sie haben zudem eine große Bedeutung in

der Energieerzeugung, wie beispielsweise in Kraftwerken, in der hydrothermalen

Behandlung von Biomasse, sowie in der hydrothermalen Vergasung, oder der

hydrothermalen Oxidation von Gefahrstoffen.

Am Paul Scherrer Institut wird ein Prozess entwickelt, Biomasse mit einem hohen

Wassergehalt in Methan als Energieträger umzuwandeln. Damit soll ein Beitrag geleistet

werden, den weltweiten Energiebedarf in Zukunft nachhaltig und CO2-neutral zu decken.

Bei diesem Prozess ist das Phasenverhalten der anorganischen Bestandteile, insbesondere

der Salze des Ausgangsmaterials, entscheidend, da diese zur Verstopfung der Anlage oder zu

Beschädigung des verwendeten Ru/C Katalysators führen können. Deshalb müssen diese

Bestandteile aus dem Prozessstrom abgetrennt werden. Die abgeschiedenen Salze können

als Dünger in der Biomassegewinnung verwendet werden, um somit den Nährstoffkreislauf

zu schließen. Dies ist ökologisch und ökonomisch sinnvoll und erhöht die Wertschöpfung des

Prozesses zur Methangewinnung.

Um den Prozess hinsichtlich der Salzabscheidung und Rückgewinnung zu optimieren, müssen

grundlegende Fragestellungen zum physikalisch-chemischen Verhalten der hydrothermalen

Salzlösungen untersucht werden. Das Verhalten der verschiedenen Typen von Salzen, deren

Wechselwirkung mit dem wässrigen Lösungsmittel und die Kontrollierbarkeit des

Phasenverhaltens sind die massgeblichen Aspekte dieser Fragestellungen.

Das Ziel dieser Dissertation ist es, das Phasenverhalten und dessen mikroskopischen

Ursprung anhand von Modelllösungen zu untersuchen und das erhaltene Wissen für die

Verbesserung der hydrothermalen Vergasung anzuwenden.

Für die Untersuchungen wurden Modellsubstanzen zweier verschiedener Gruppen von

Salzen verwendet. So genannte Typ 1 Salze besitzen kontinuierliche Löslichkeitsgrenzen in

ihren wässrigen Lösungen, die vom Tripelpunkt des reinen Wassers bis zum Tripelpunkt des

Page 14: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

X

reinen Salzes reichen. Die Phasendiagramme dieser Salze werden oft durch das Auftreten

einer bestimmten Art von Flüssig-Entmischung verkompliziert (sogenannter Typ 1d). Die

zweite Gruppe von Salzen (Typ 2 genannt) zeigt eine diskontinuierliche Löslichkeitslinie.

Schnittpunkte der kritischen Linien (Gas-Flüssig) mit der Löslichkeitslinie führt zu einem

Bereich zwischen den kritischen Endpunkten p und Q, in dem ein einphasiges Fluid existiert,

das typischerweise wenig Salz enthält und im Gleichgewicht mit dem festen Salz steht. In

einer früheren Arbeit konnten die Typ 1d Salze effizient aus dem Prozessstrom der

hydrothermalen Vergasung abgeschieden und zurückgewonnen werden, wohingegen die

Typ 2 Salze zu Verstopfung der Anlage aufgrund des Ausfallens von festem Salz führten.

Der erste Teil dieser Dissertation beschreibt, wie mit Hilfe der dynamischen

Differenzkalorimetrie (DSC) das Phasenverhalten von binären und ternären Salzlösungen

untersucht wird. Auf Basis der binären Salzlösungen konnte gezeigt werden, dass die DSC-

Messungen sehr gut dazu geeignet sind die Art der Phasenübergänge in hydrothermalen

Salzlösungen zu untersuchen. So können das Ausfallen von festem Salz, die Bildung einer

Flüssig-Entmischung und Phänomene, die zur Homogenisierung von Gas- und Flüssigphase

führen, unterschieden werden.

Die gewonnenen Erkenntnisse wurden dazu genutzt, ternäre Salzlösungen, die aus zwei

Salzen und Wasser als Lösungsmittel bestehen, zu untersuchen. In seltenen Fällen kann die

Mischung von zwei Typ 2 Salzen, in Abhängigkeit von Salzkonzentration und

Molzahlverhältnis der beiden Salze, das Phasenverhalten einer Typ 1d Lösung zeigen. Dieser

Effekt wird dadurch erreicht, dass eine Flüssig-Entmischung stabilisiert wird, die in den

binären Salzlösungen metastabil ist. Dies ist zum Beispiel im System NaxK2-xSO4-H2O der Fall.

Ternäre Lösungen von Natrium- oder Kaliumsulfat mit Magnesiumsulfat zeigten hingegen

keinerlei Stabilisierung der Flüssig-Entmischung. Bei der Verwendung von Carbonat als

Anion, bildet sich in den Lösungen mit Natrium und Kalium als Kation nur dann ein Typ 1(d)

Phasenverhalten aus, wenn das Stoffmengenverhältnis von Na+ zu K+ kleiner als 10:1 war.

Gemischte Lösungen von Sulfaten und Hydrogensulfaten bei denen alternativ Natrium oder

Kalium als Gegenionen verwendet wurden führten zu grossen Zusammensetzungsbereichen

in denen die Flüssig-Entmischung stabil ist, bis hin zu SO42-:HPO4

2- Stoffmengenverhältnissen

von 1:10. Diese Ergebnisse zeigen, dass zur bevorzugten Beeinflussung des

Phasenverhaltens, also zum Erreichen eines Typ 1d Verhaltens, eine Zugabe von K2HPO4 von

Page 15: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XI

Vorteil ist. Weiterhin konnte gezeigt werden, dass das Phasenverhalten in ternären

Lösungen den gleichen Tendenzen folgt, wie in den binären Mischungen. Dies bedeutet, dass

Lösungen von kleinen, stark geladenen Ionen eher ein Typ 2 Phasenverhalten aufweisen als

Lösungen von weniger geladenen, grösseren Ionen, welche Typ 1d Verhalten hervorrufen.

Zusätzlich scheint bei polyatomaren Anionen die Anzahl von Koordinationsstellen eine Rolle

zu spielen.

Im zweiten Teil dieser Dissertation wird die mikroskopische Struktur der hydrothermalen

Salzlösungen untersucht. Mittels FT-IR Raman Spektroskopie wurden hydrothermale

Natriumsulfat-Lösungen (Typ 2) untersucht. Das Raman-Signal, das unter

Umgebungsbedingungen dem freien Sulfat-Ion zugeordnet wird, zeigte hierbei eine

Verschiebung bei zunehmender Temperatur hin zu niedrigeren Energien. Ein weiteres

Raman-Signal erscheint, bei Annäherung von Druck- und Temperatur an die

Löslichkeitsgrenze, welches an Intensität zunimmt, je näher die Versuchsbedingungen dieser

Grenze kommen. Mit Hilfe von molekulardynamischer Simulationen (MD) konnte dieses

Raman-Signal ionischen Clustern zugeordnet werden, die aus bis zu ~5 Sulfat-Ionen

entstehen und als Vorläufer des festen Salzes in Erscheinung treten.

Wurden die Natrium-Ionen teilweise durch Kalium-Ionen ersetzt, verschwand das Raman-

Signal der beschriebenen Cluster. DSC-Messungen offenbarten einen starken

Löslichkeitsanstieg. Weitere MD-Simulationen indizierten, dass der Austausch von Natrium-

Ionen durch Kalium-Ionen dazu führt, dass grössere Cluster zerfallen, da die elektrostatische

Wechselwirkung der Kalium-Ionen mit den Sulfat-Ionen geringer ist. Dies ist auf den

grösseren Ionenradius der Kalium-, gegenüber den Natrium-Ionen, zurückzuführen.

Weiterhin zeigten sich ionische, polymere Kettenstrukturen in den MD Simulationen der

ternären Systeme. In der Literatur wurde bereits früher vermutet, dass solche Strukturen für

die Ausbildung von Flüssig-Entmischungen verantwortlich sind. Unsere Ergebnisse stützen

diese These, da eine Flüssig-Entmischung auch im System NaxK2-xSO4-H2O präsent ist, welche

in den jeweiligen binären Systemen nicht vorhanden ist.

Der dritte Teil dieser Dissertation befasst sich mit einem Entwurf eines Salzabscheiders, der

die Erkenntnisse zum Phasenverhalten der Salzlösungen nutzt. Anstatt des ursprünglich in

der hydrothermalen Vergasung am PSI verwendeten Design, bei dem die Biomasse an der

Page 16: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XII

Oberseite des Reaktors mittels eines Tauchrohrs zugeführt wird, wird nun ein Steigrohr

benutzt, welches die Zuführung von unten in den Salzabscheider gewährleistet. Dies führt

zur Ausbildung einer Pseudo-Grenzfläche zwischen einem Fluid niedriger Dichte im oberen

Teil des Reaktors und einem Fluid hoher Dichte im unteren Bereich. In der weniger dichten

Phase besitzen die Salze eine geringe Löslichkeit. Diese Mischung aus Wasser und

organischen Bestandteilen kann zur Methanierung mittels eines Katalysators abgeführt

werden. Das dichte, kältere Fluid hält die Salze in Lösung und kann in Form einer

konzentrierten Lösung am Boden des Reaktors separiert werden. Erste Versuche mit dem

neuen Entwurf lieferten vielversprechende Ergebnisse, sowohl für Modelllösungen, als auch

für echte Biomasse.

Zusammenfassend lässt sich feststellen, dass sich das Phasenverhalten von Mischungen

durch Zugabe eines passenden Salzes erfolgreich verändern lässt. Basierend auf diesem

Wissen wurde ein neuer Salzabscheider entworfen. Weiterhin wurde die Tendenz zur

Ausbildung von Clustern, welche überwiegend durch den Ionenradius und die Ladung der

Ionen bestimmt wird, als entscheidendes Kriterium für die mikroskopischen Ursachen des

makroskopischen Phasenverhaltens identifiziert.

Page 17: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XIII

Table of Contents

Danksagung .................................................................................................................................. III

Abstract .......................................................................................................................................... V

Zusammenfassung ........................................................................................................................ IX

Table of Contents ........................................................................................................................ XIII

Introduction ............................................................................................................. 1 Chapter 1

1.1. Background and Motivation ........................................................................................ 1

1.2. PSI’s Hydrothermal Gasification Process .................................................................... 3

1.3. Inorganic Constituents in Biomass .............................................................................. 4

1.4. Aim of the Project ....................................................................................................... 5

Literature Survey and Theoretical Background ....................................................... 7 Chapter 2

2.1. Water at Elevated Temperature and Pressure ........................................................... 7

2.2. Aqueous Salt Solutions at Elevated Temperature and Pressure ................................ 9

2.3. Phase Behavior of Binary Aqueous Salt Solutions..................................................... 11

2.4. Phase Behavior of Ternary Aqueous Salt Solutions .................................................. 17

2.5. Salt Formation and Nucleation ................................................................................. 21

2.6. Speciation and Microscopic Structure of Salt Solutions ........................................... 23

Part A Macroscopic Phase Behavior of Hydrothermal Salt Solutions ......................................... 27

Studies on the Phase Behavior of Salt Solutions Using High Pressure Chapter 3

Differential Scanning Calorimetry ............................................................................................... 29

3.1. Introduction ............................................................................................................... 29

3.2. Methods: HP-DSC ...................................................................................................... 30

3.3. Results and Discussion: HP-DSC ................................................................................ 33

Page 18: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XIV

3.3.1. Binary Solutions ..................................................................................................... 33

3.3.1.1. Na2SO4-H2O ............................................................................................................ 33

3.3.1.2. K2SO4-H2O .............................................................................................................. 36

3.3.1.1. K2HPO4-H2O............................................................................................................ 37

3.3.2. Ternary Solutions ................................................................................................... 38

3.3.2.1. Na2SO4-K2SO4-H2O ................................................................................................. 41

3.3.2.2. Na2SO4-Na2CO3-H2O ............................................................................................... 43

3.3.2.3. Na2SO4-MgSO4-H2O ............................................................................................... 45

3.3.2.4. Na2SO4-Na2HPO4-H2O ............................................................................................ 46

3.3.2.5. K2CO3-Na2CO3-H2O ................................................................................................. 47

3.3.2.6. K2SO4-MgSO4-H2O .................................................................................................. 48

3.3.2.7. K2SO4-K2HPO4-H2O ................................................................................................. 49

3.3.3. Influence of Organics ............................................................................................. 49

3.4. Conclusions and Outlook ........................................................................................... 50

Studies on the Phase Behavior of Salt Solutions with the Optical Cell ................. 53 Chapter 4

4.1. Introduction ............................................................................................................... 53

4.2. Methods: Optical Cell ................................................................................................ 53

4.3. Results and Discussion: Optical Cell .......................................................................... 55

4.3.1. K2SO4-H2O .............................................................................................................. 55

4.3.2. NaxK2-xSO4-H2O ....................................................................................................... 57

4.4. Conclusions and Outlook ........................................................................................... 59

Part B Microscopic Structure of Hydrothermal Salt Solutions .................................................... 61

Page 19: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XV

Studies on the Speciation in Hydrothermal Salt Solutions with X-ray Chapter 5

Absorption Spectroscopy............................................................................................................. 63

5.1. Introduction ............................................................................................................... 63

5.2. Setup and operation description............................................................................... 64

5.3. Improvements on Cell and Windows ........................................................................ 67

5.3.1. Windows ................................................................................................................ 67

5.3.2. Distance Rings ........................................................................................................ 71

5.3.3. Gaskets................................................................................................................... 72

5.4. Conclusion and Recommendations ........................................................................... 73

Studies on the Speciation in Hydrothermal Salt Solutions with Chapter 6

FTIR-Raman Spectroscopy ........................................................................................................... 75

6.1. Introduction ............................................................................................................... 75

6.2. Methods: FTIR-Raman Spectroscopy ........................................................................ 76

6.3. Results and Discussion: FTIR-Raman Spectroscopy .................................................. 80

6.3.1. Na2SO4-H2O ............................................................................................................ 80

6.3.2. Li2SO4-H2O .............................................................................................................. 85

6.3.1. Na2CO3-H2O............................................................................................................ 86

6.3.2. Na2HPO4-H2O ......................................................................................................... 87

6.3.3. Na2SO4-K2SO4-H2O ................................................................................................. 88

6.4. Conclusions and Outlook ........................................................................................... 90

Studies on the Speciation in Hydrothermal Salt Solutions with Chapter 7

Molecular Dynamics Simulations ................................................................................................ 93

7.1. Introduction ............................................................................................................... 93

Page 20: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

XVI

7.2. Methods: MD-Simulations ........................................................................................ 94

7.3. The Cluster Analysis Program .................................................................................... 97

7.4. Results and Discussion: MD-Simulations .................................................................. 98

7.4.1. Simulations of the System Na2SO4-H2O ................................................................. 98

7.4.2. Simulations of the System NaxK2-xSO4-H2O .......................................................... 111

7.5. Conclusions and Outlook ......................................................................................... 122

Part C Design of a Salt Separator for the Hydrothermal Gasification Of Biomass .................... 127

Salt Separator Design .......................................................................................... 129 Chapter 8

8.1. Introduction ............................................................................................................. 129

8.2. Theoretical Considerations/Patent Application ...................................................... 130

8.3. Results ..................................................................................................................... 133

8.4. Summary and Outlook ............................................................................................ 135

Part D Conclusions ..................................................................................................................... 137

Summary, Conclusions and Prospects ................................................................. 139 Chapter 9

9.1. Summary and Conclusions ...................................................................................... 139

9.2. Recommendations for Future Research ................................................................. 142

References ................................................................................................................................. 145

Curriculum Vitae ........................................................................................................................ 158

Publications Related to this Thesis ............................................................................................ 159

Page 21: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 1

1

Introduction Chapter 1

1.1. Background and Motivation

With a growing world population and ongoing technological progress, the demand for

energy increases tremendously. To date, this demand is mostly met by fossil fuels such as

coal, lignite, oil and natural gas. As those fossil fuels are not of infinite supply, the need to

substitute them with “renewable” energy sources to cover future energy demand is

increasingly urgent. Furthermore, combustion of fossil fuels has led to a dramatic increase in

the carbon dioxide content of Earth's atmosphere, which triggers the climate change caused

by the global warming. In order to reach the climate goals, i.e. a reduction of CO2 emissions

and limiting the global warming to a certain extent, a substitution of fossil fuels with

renewable energy sources is absolutely essential.

The World Energy Outlook 20121 predicts that the share world primary energy demand met

by fossil fuel combustion will decrease from 81% in 2010 to 75-63% in 2035, whereas the

usage of renewable energy increases from 13% to 18-27% over the same time period

(depending on the scenario). The renewable energies mainly comprise traditional biomass

(i.e. burning of wood), hydropower, wind, and solar energy and to a lower extent the usage

of “modern” biomass (i.e. Synthetic Natural Gas, or liquid fuels from biomass). Geothermal

and tidal energy usage is very limited, due to the special conditions needed for exploiting

these sources. Wind and solar power suffer from strong fluctuations (over days and seasons)

and, in combination with the present lack of electricity storage, this leads to problems for

the power grid and hence to a relatively low percentage of renewable energy in the overall

energy consumption.

Using biomass as an energy carrier overcomes the storage problem. Whereas the usage of

wood is well established, wet biomass usage (sewage sludge, manure, organic waste etc.)

has developed only over the past 25, with an increasing number of anaerobic digestion

plants. Those plants produce so-called "biogas," which is usually used for electricity and heat

production directly on-site. The drawbacks of anaerobic digestion are the low efficiency

Page 22: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

1.1 Background and Motivation

2

(~25%), low space-time yields and an incomplete digestion, leading to residues with still high

energy content, which may lead to methane emissions to the atmosphere.

To improve the possibilities of using wet biomass for energy production, a thermochemical

process was developed at the Paul Scherrer Institute,2 which makes use of the unique

properties of water at elevated temperatures and pressures to produce synthetic natural gas

(SNG) at high space-time yields and high efficiencies. One crucial step in this process is the

separation of the inorganic constituents of the biomass from the process stream in order to

avoid blocking of the plant and poisoning of the catalyst which is used. To optimize this

separation, a detailed insight on the phase behavior and speciation of biomass-related

aqueous salt solutions at high temperatures and pressures is needed. The aim of the present

thesis is to investigate such salt solutions at the given conditions and deepen the

understanding of their behavior in order to be able to optimize the separation process.

Page 23: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 1

3

1.2. PSI’s Hydrothermal Gasification Process

At the Paul Scherrer Institute, a process for the catalytic supercritical water gasification

(SCWG) was developed2, 3 in order to provide faster and more efficient utilization of the

energy content of wet biomass. The term "supercritical" here refers to conditions above the

critical point of pure water (374°C, 22.1 MPa). The original configuration (Figure 1-1) consists

of:

1. A feed tank for storing the biomass slurry.

2. A high-pressure pump/slurry feeder to pressurize the feed.

3. A reactor for preheating the feed to 300-370°C, leading to cracking of the biomass

molecules into smaller parts.

4. The salt separation vessel, where the feed is injected through a dip tube into a hot

zone (420-450°C), salts are removed as a concentrated brine at the bottom and the

water/biomass mixture exits radially at the top of the vessel.

5. The catalytic reactor, where the organic molecules of the biomass are converted to

methane and CO2 at 400-450°C using a Ru/C catalyst.

6. A phase separation unit to control the pressure and separate the gas from the water.

Page 24: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

1.3 Inorganic Constituents in Biomass

4

Figure 1-1: Outline of PSI's hydrothermal gasification process. Taken from F. Vogel.2 Used with permission.

1.3. Inorganic Constituents in Biomass

Phosphorus occurs in biomass typically in form of phospholipids as the main building blocks

of cell membranes, as well as in form of adenosine phosphates, which act as the ‘fuel’ of the

metabolism. It also forms parts of the DNA. Sodium and potassium are present in biomass as

they are the most important ions for the osmotic pressure inside the cells as well as for

membrane polarization of the cells in animals. Sulfur is usually bound as organic sulfur in

proteins, which contain the amino acids cysteine and methionine. The formation of disulfide

bridges between two cysteine units in proteins is necessary for the correct folding of the

proteins and hence their biological function.

Biomass slurry

Water

CO2

SNG(to pipeline, gas engine, fuel

cell, gas turbine)High pressure slurry

pump

Preheater

(heat

recovery)

Gas fired burner

Catalytic reactor

(gasification &

methanation)

Cooler

Phase

separator

Air

PWS

Superheater &

salt separator

Salt brine

350-380°C

420-450°C

400°C150°C

300 bar

300 bar

Page 25: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 1

5

1.4. Aim of the Project

This thesis is divided into three major parts. The first part deals with the macroscopic phase

behavior of binary and ternary aqueous salt solutions at hydrothermal conditions. In the

second part the microscopic structure of such solutions was investigated. The third part

closes the gap towards the application of the knowledge gathered in the first two parts.

The tasks in detail are listed as follows:

Developing and applying a method for determining phase behavior of hydrothermal

salt solution using high pressure differential scanning calorimetry (Part A, Chapter 3)

Studying the nature of phase transitions which occur in such salt solutions by using

an optical cell (Part A, Chapter 4)

Attempts to study the speciation in sulfate solutions using X-ray absorption

spectroscopy (Part B, Chapter 5)

Investigation of speciation in salt solutions close to precipitation conditions using

FT-IR-Raman-Spectroscopy (Part B, Chapter 6)

Molecular dynamics simulations on the ion association and cluster formation at high

temperatures in Na2SO4 solutions (Part B, Chapter 7)

Suggestions leading to a new design of the salt separation unit in PSI’ hydrothermal

gasification process (Part C, Chapter 8)

An overall conclusion will be made in Part D (Chapter 9), complemented with suggestions for

further research in this field.

Page 26: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

1.4 Aim of the Project

6

Page 27: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

7

Literature Survey and Theoretical Background Chapter 2

2.1. Water at Elevated Temperature and Pressure

Water possesses unique properties, such as a density maximum at 4°C, a high boiling point

of 100 °C at atmospheric pressure, a very high static permittivity of 80 (20 °C, 1 atm), and

very high pressure and temperature of the critical point (374 °C, 22.1 MPa). At elevated

pressures and temperatures, the static permittivity decreases to values typical for organic

solvents. This is also reflected in single phase equilibria with organic solvents e.g. in the

benzene-water system at high pressures and temperatures.4 The density decreases by

approximately 60% to 322 kg/m3 at the critical point.

Figure 2-1: T-ρ diagram of the gas-liquid coexistence curve of pure water (blue line) above 25 °C; data from Ref. 5.

0

50

100

150

200

250

300

350

400

0 200 400 600 800 1000

T /°

C

ρ / kg/m³

gas + liquid

gas liquid

supercritical fluid

critical point

Page 28: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.1 Water at Elevated Temperature and Pressure

8

Figure 2-2: Detail of the p-T diagram of pure water between 100 and 450 °C; gas-liquid coexistence line in blue, critical isochore in green; data from Ref. 5.

Figure 2-3: Static permittivity and ionization product of water at 30 MPa, calculated from Ref. 6 and Ref. 7.

0

5

10

15

20

25

30

35

40

45

100 150 200 250 300 350 400 450

p /

MP

a

T / °C

liquid

gas

critical point

supercritical fluid

-22

-21

-20

-19

-18

-17

-16

-15

-14

-13

-12

-11

-10

0

10

20

30

40

50

60

70

80

0 50 100 150 200 250 300 350 400 450 500

log 1

0 K

W

ε 0

T / °C

Page 29: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

9

2.2. Aqueous Salt Solutions at Elevated Temperature and Pressure

In nature, water essentially never occurs as a truly "pure" substance. Even when vaporized

by sunlight at Earth's surface and condensed or frozen into clouds or fog, the water

comprises impurities from the atmosphere, which results in a mixture of different

components with physical properties differing from the ones of the pure substance. In

Earth’s crust, where water interacts with rocks, minerals and magmas, the water takes up

electrolytes until solubility equilibria are reached. The depth in the crust at which water is

situated also determines its pressure (hydrostatic or lithostatic, according to the overlying

load), and its temperature is commonly controlled by the geothermal gradient (usually

around 25°C/km depth). Aqueous salt solutions play a key role in heat transport in the

Earth’s crust as well as in the transport of metals, including the formation of economic ore

deposits.8

Hydrothermal salt solutions also play a role in energy production. The addition of salts to

process water in conventional and nuclear power plants was discussed in the context of

prevention of corrosion.9 In addition, naturally heated, low- to moderately-saline geo-fluids

(analogous to hot springs) are used in geothermal power plants, where scaling issues are

critical for the process.

Systematic basic research on hydrothermal salt solutions started in the group of Prof. E.U.

Franck in the 1950s (at that time in Göttingen, later in Karlsruhe), in W.L. Marshall’s group at

Oak Ridge National Laboratory and in Russia in the group of M.I. Ravich. The two most

studied systems are NaCl-H2O and Na2SO4-H2O, most likely due to their high abundance in

nature and their importance in geological processes.

The latest full summary on high-pressure, high-temperature research was presented by

Valyashko10 in 2008.

A remark on the terminology in this thesis: The terms “gas” and “vapor” are used

synonymously. The term “critical point” for a pure substance refers to the endpoint of the

coexistence line of two phases, where the two phases become equal in their physical

properties. Mostly this term is used for gas-liquid criticality (G=L), also referred to as

vapor-liquid criticality (V=L); nevertheless it is also valid for other critical phenomena, which

Page 30: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.2 Aqueous Salt Solutions at Elevated Temperature and Pressure

10

happen e.g. in case of liquid immiscibilities. In binary mixtures the critical points in the phase

diagrams of the pure substances transfer into critical lines spreading into the binary space.

Hence, a “supercritical” binary fluid would be a homogeneous fluid with pressure and

temperature above the actual critical parameters at the respective concentration. Since

those parameters vary with the concentration (and since in binary systems liquid-gas

immiscibility can extend to pressures and temperatures greater than the respective point on

the critical line) it is more useful to speak about a homogeneous fluid instead of a

supercritical one, and leaving the term critical for endpoints of coexistence lines of different

phases.

Page 31: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

11

2.3. Phase Behavior of Binary Aqueous Salt Solutions

Scott and Van Konyenburg11 used the Van der Waals equation to create a set of seven

different phase diagrams for fluid phase equilibria, not including solid phases. Later

Valyashko12, 13 derived 26 full phase diagrams (which also include fluid-solid equilibria) for

binary systems using topological methods. The main division between the types of phase

diagrams is the one between type 1 and type 2. The type 1 phase diagrams have a

continuous three phase line, meaning that critical phenomena are absent in presence of a

solid phase, whereas in the type 2 phase diagrams the three phase line intersects the critical

line, leading to two separate three phase regions which are divided by a region, where only a

single fluid or a fluid-plus-solid are present.

Figure 2-4: p-T projections of type 1 (left) and type 2 (right) phase diagrams. Thin solid lines: two phase equilibria in the pure substances; thick solid lines: three phase equilibria in the binary mixture; dot-dashed line: critical line in the binary mixture; TPW, TPS: triple points of water and salt, respectively; E: eutectic points; CPW, CPS: critical points of water and salt, respectively; p, Q: lower and upper critical endpoint (G=L-S, L1=L2-S), respectively. Adapted from Valyashko.12

Experimentally, the most important ones are the so called type 1a (e.g. NaCl-H2O, KCl-H2O,

CaCl2-H2O), type 1d (e.g. K2HPO4-H2O, Na2HPO4-H2O, K2CO3-H2O, Na2B4O7-H2O) and type 2d’

(e.g. Na2SO4-H2O, K2SO4-H2O, MgSO4-H2O, Na2CO3-H2O, SiO2-H2O), which will be explained in

detail below.

TPs

p

T

TPWCPs

CPW

ETPs

p

T

TPW CPs

CPW

E

pQ

Page 32: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.3 Phase Behavior of Binary Aqueous Salt Solutions

12

The type 1a phase diagram (Figure 2-5) exhibits a continuous three phase line (Gas-Liquid-

Solid, G-L-S) from the triple point of pure water to the triple point of the pure salt. That

means that the solubility of the salt is constantly increasing with increasing temperature or

in other words: The temperature coefficient of solubility (t.c.s.) is positive. Furthermore, the

critical line, leading from the critical point of pure water to the critical point of the pure salt

shows no discontinuity. At pressures above that line, only one single fluid phase is present.

Figure 2-5: T-x, p-T and p-x projections of a type 1a phase diagram. Labeling see Figure 2-4. Adapted from Valyashko.12

Type 1a

TPs

p

T

TPWCPs

CPW

E

TPW

p

xW S

CPW CPS

TPSE

EV

CPs

CPW

TPW

EL

sat. G

sat. L

x

T

W S

Page 33: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

13

Figure 2-6: Phase diagram of the system NaCl-H2O to 1000°C and 2200 bar. H: halite (solid NaCl), V: vapor (gas) phase, L: liquid phase. Taken from Driesner.14 Used with permission.

Phase diagrams of the so called type 1d (Figure 2-7) also have a continuous three phase line

(G-L-S), but they also exhibit a liquid immiscibility, where a dilute liquid phase (L1) is in

equilibrium with a concentrated liquid phase (L2). The immiscibility is characterized by a

three phase line (G-L1-L2) and starts at a lower critical point N, where the two liquid phases

become equal in terms of composition, and an upper critical point R, where the gas phase

and the dilute liquid phase become equal. The critical line hence leads from the critical point

of the pure water to the upper critical point R of the liquid immiscibility and from the lower

critical point N to the critical point of the pure salt.

Page 34: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.3 Phase Behavior of Binary Aqueous Salt Solutions

14

Figure 2-7: T-x, p-T and p-x projections of a type 1d phase diagram. Labeling see Figure 2-4. N: lower critical point V-L1=L2; R: upper critical point V=L1-L2. Adapted from Valyashko.12

The type 2d’ phase diagram (Figure 2-8) exhibits a discontinuous three phase (G-L-S) line,

which intersects the critical line at the lower critical endpoint p and the upper critical

endpoint Q. The three phase (G-L-S) line, i.e. the solubility line, is characterized by a change

in the sign of the t.c.s., meaning that at a certain temperature the solubility of the salt starts

to decrease. That leads to the intersection of the G-L-S line with the critical line. Hence G-L

critical phenomena are present in equilibrium with a solid phase. Typically, the p-T

conditions of the lower critical endpoint p are in the proximity of the critical point of pure

water (374°C, 22.1 MPa) and the solute content is very low. In contrast, the upper critical

endpoint is at fairly high pressures and temperatures and high solute concentrations (e.g. in

SiO2-H2O: pQ = 9.7 kbar TQ = 1080°C x(SiO2) ≈ 75 wt.%15). Note that in the upper three-phase

region, at conditions close to the upper critical endpoint Q, the dilute phase is more liquid-

like rather than gaseous,12 and then gradually becomes more gas-like towards the triple

point of the pure salt.

Type 1d

TPs

p

T

TPWCPs

CPW

E

R

N

R

EG

CPs

TPs

CPW

TPW

EL

sat. G

sat. L

T

W S

R

N

x

p

xW S

E

TPW

CPW

CPS

TPS

N

R R

Page 35: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

15

Figure 2-8: T-x, p-T and p-x projections of a type 2d' phase diagram. Labeling see Figure 2-4. Adapted from Valyashko.12

Type 2d‘

TPs

p

T

TPW CPs

CPW

E

p

Q

p

xW S

E

TPW

CPW

CPS

TPS

p

Q

EG

p

Q

CPs

TPs

CPW

TPW

sat. G

sat. L

T

W S

EL

x

Page 36: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.3 Phase Behavior of Binary Aqueous Salt Solutions

16

Figure 2-9: Sketch of the lower three phase region of a type 2 phase diagram. In light blue are isothermal cuts through the G-L space.

Page 37: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

17

2.4. Phase Behavior of Ternary Aqueous Salt Solutions

Valyashko derived and classified ternary phase diagrams from the above mentioned binary

ones. In general, ternary phase diagrams with one volatile compound (pure water in the

present case) and two non-volatile compounds (two different salts) are determined by their

binary boundary phase diagrams,12 which were described above.

With changing the molar fraction of one salt from 0 to 1, relative to the second salt, phase

behavior will change gradually from the behavior of the binary subsystem of the second salt

to the one of which the concentration is increased. Note that all ternary phase diagrams

mentioned and described here are in principal only valid for systems in which the binary

phase diagram for the mixture of the two salts is of type 1a, meaning that it exhibits

continuous solid and fluid solutions of the two salts.12 More complicated phase diagrams for

mixtures of the two salts will most likely also complicate the phase diagrams of the ternary

solutions.

Due to the increased number of degrees of freedom, the critical endpoints of the binary

subsystems will be represented as critical lines in the ternary phase diagrams.

The transformation from one boundary type to another can lead to multiple critical

endpoints such as the so called pQ (G=L-S) critical endpoint (in case one or both of the

boundary subsystems is of type 2), where type 2 behavior turns into type 1 behavior,12 or

the NR (L1=L2=G) critical endpoint, where a liquid immiscibility of the d-type starts/ends (in

case one or both of the subsystems have a d-type immiscibility).

The ternary systems of interest in this thesis consist of subsystems of either type 1d

(K2CO3-H2O, Na2HPO4-H2O, K2HPO4-H2O) or type 2d’ (Na2SO4-H2O, K2SO4-H2O, Na2CO3-H2O,

MgSO4-H2O), hence only the relevant phase diagrams will be discussed here.

The ternary phase diagrams are in general four dimensional, hence it is impossible to

illustrate them, without using some kind of projection. A useful representation to describe

the full phase diagrams is the T-X* projection, where X*=XB/(XB+XC).12, 16 Here the subscripts

B and C denote the two salts, subscript A would be the solvent, i.e. water in our case.

Page 38: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.4 Phase Behavior of Ternary Aqueous Salt Solutions

18

According to Valyashko’s classification,12, 16 ternary systems with two type 1d subsystems

can lead to ternary phase diagrams denoted IIIα, IIIβ, IIIγ and IIIδ.

In the IIIα system both liquid immiscibilities emerge into the ternary space and both of them

end in tricritical endpoints NR (G=L1=L2).

IIIβ systems exhibit a liquid immiscibility over the whole range of X*.

The more exotic systems IIIγ and IIIδ show another phenomenon, which leads to a splitting

of the liquid immiscibility region into two parts between two ternary critical points both

called N’N, in the IIIγ case. In the IIIδ system the upper part of the region, where the splitting

of the liquid immiscibility is present, intersects with the L1=G-L2 critical line, leading to two

critical endpoints of the NR type (G=L1=L2).

Figure 2-10: T-X* projection of ternary phase diagrams with two type 1d subsystems. Adapted from Valyashko.12

Valyashko12, 16 also presents ternary phase diagrams with one subsystem of type 1d and one

subsystem of type 2d’. He suggests four different versions designated IIIα-1, IIIα-2, IIIα-3 and

IIIβ-3.

In the IIIα-1 systems both liquid immiscibilities (the stable one from the 1d subsystem and

the metastable one from the 2d’ subsystem) end in ternary critical points of the NR type

(G=L1=L2).

Page 39: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

19

Figure 2-11: T-X* projections of ternary phase diagrams with a binary subsystem of type 1d (left side) and a subsystem of type 2d’ (right side). Adapted from Valyashko.12

The IIIα-2 type exhibits a transformation of the metastable liquid immiscibility into a stable

one, around the temperatures of the lower critical endpoint q (G=L-S) of the 2d’ subsystem.

For the IIIα-3 systems this transformation happens above the temperatures of the upper

critical endpoint Q (L1=L2-S) of the binary 2d’ subsystem.

All IIIα systems have in common that the stable liquid immiscibility emerging from the 1d

subsystem and the metastable liquid immiscibility, which arises through the 2d’ subsystem

are divided by a region of liquid miscibility.

In contrast to that, in the IIIβ-3 phase diagrams the above mentioned miscibility region is

absent and the stable liquid immiscibility transforms into the metastable one with increasing

X*. This transformation takes place, where the critical lines emerging from the critical

endpoints p and Q interfere with the critical line connecting N-Nms (L1=L2-G) and R-Rms

(L1=G-L2).

Page 40: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.4 Phase Behavior of Ternary Aqueous Salt Solutions

20

For ternary systems, where both subsystems are of type 2d’, Valyashko12, 16 suggests two

types of phase diagrams, denominated IIIβ-4 and IIIβ-6.

In the IIIβ-6 phase diagram the G=L-S and L1=L2-S critical lines do not intersect the

metastable L1=G-L2 and G-L1-L2 critical lines, respectively. Hence the liquid immiscibility is

metastable over the whole range of X*.

In contrast to the IIIβ-6 phase diagram, the IIIβ-4 system exhibits interference of the G=L-S

and L1=L2-S critical lines with the L1=G-L2 and G-L1-L2 critical lines, leading to a stabilization

of the liquid immiscibility in the region between the two ternary critical points of type LN

(L1=L2-G-S) and the two ternary critical points of type pR (L1=G-L2-S).

Figure 2-12: T-X* projections of ternary phase diagrams with two subsystems of type 2d’. Adapted from Valyashko.12

Many ternary phase diagrams have been investigated in the group of Valyashko, mainly at

concentrations above 5 wt.% and at temperatures above 390 °C.17-24

Page 41: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

21

2.5. Salt Formation and Nucleation

Precipitation of a solid from a solution is classically described by the process of Ostwald

ripening.25-27 In this theory, small crystallites form from a supersaturated solution. Some of

those crystallites grow at the expense of others. Particles below a critical radius dissolve

readily. Gebauer and Cölfen28 suggested that precipitation of a solid (in this case CaCO3) is

preceded by formation of amorphous clusters of ions (termed prenucleation clusters), which

subsequently transform into the stable crystalline phase. This new approach is also called

non-classical nucleation. Both theories are mainly applied to precipitation at room

temperature; whether they apply also to precipitation of solid salt from hydrothermal

solutions is not yet known.

Early studies on salt separation and precipitation were related to the supercritical water

oxidation (SCWO) process and dealt with the solubility of salts in water above the critical

temperature and pressure of pure water, as well as with crystallization processes occurring

there. In the SCWO process, salts are formed during the oxidation of organic matter and

other salts can be added to the process stream for adjusting the pH-value of the reaction

mixture. Previous studies have used different approaches e.g. precipitation on a “hot finger”,

special reactor designs (e.g. MODAR reactor, transpiring wall reactor).29-38 The results of

these studies are commonly affected by the setup design and restricted to a limited range of

parameters (pTX).

For a successful operation of SCWO and SCWG processes also the nucleation and

precipitation kinetics of salts at and above the critical point of pure water are important.

Armellini et al.36-38 studied shock crystallization by injecting a salt solution jet into

supercritical water. They found that the crystallite size can be modeled by diffusion-limited

growth, but agglomeration could not sufficiently be described by simple particle collision

models. Hodes et al.39 found precipitation rates of around 1 g(salt)/min when precipitating

sulfate salts from a solution at a “hot finger”. The deposition rates where in accordance with

a simple heat and mass transfer model. Rogak and Teshima40 investigated salt deposition in

a tubular reactor and modeled the deposition with a diffusion-limited approach. They

concluded that precise modeling of salt precipitation at conditions around the critical point

of pure water is hindered by a lack of information on the properties of the salt solutions at

Page 42: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.5 Salt Formation and Nucleation

22

such conditions. As experimental measurements of nucleation rates are complicated,

recently Molecular Dynamics (MD) simulations have been used to study the rates of

formation and growth of ionic clusters/crystallites in NaCl41, 42 and SrCl243 solutions at

hydrothermal conditions.

A study on salt precipitation and scaling prevention in a flow reactor using ternary salt

solutions was carried out by Makaev et al.,44 who used also a vertical reactor, but with the

feed inlet at the bottom of the vessel and just one outlet on top. The focus of the study was

on influencing the phase behavior of type 2 salts by adding type 1 salts to avoid formation of

a solid salt phase in the reactor.

Over the past decades several types of reactors for the SCWO process have been developed

in order to overcome the problems of salt scaling.45, 46

In earlier investigations in the Catalytic Process Engineering group at PSI,47-49 continuous salt

separation from hydrothermal salt solutions was investigated. Binary type 1 salt solutions

showed high separation efficiencies, whereas type 2 solutions showed low separation

efficiencies and caused deposition of solid salt in the apparatus. Interestingly, a ternary

solution of the two type 2 salts Na2SO4 and K2SO4 could also be recovered efficiently. The

salt deposition in the salt separation unit was also studied with neutron radiography on the

macro-scale.50-52

Page 43: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

23

2.6. Speciation and Microscopic Structure of Salt Solutions

The microscopic structure (e.g., solvation, speciation) of salt solutions is important for

predicting phase behavior because, as we know from statistical thermodynamics, the

microscopic structure determines the macroscopic thermodynamic properties of matter.

Generally, the speciation in an aqueous electrolyte solution is dependent on three different

interactions, namely the ion-ion interaction, the ion-water interaction and the water-water

interaction. The interactions consist of electrostatic and dispersive contributions. The

interplay of the interactions will determine if a salt is dissolved, if and how the ions

associate, and/or if a salt precipitates.

Historically, soluble salts as strong electrolytes are considered to be fully dissociated in their

aqueous solutions.53, 54 Later, Bjerrum described a theory also taking the ion-association of

strong electrolytes into account.55 Eigen and Tamm56, 57 proposed, based on ultrasonic-

relaxation measurements, that the formation of ion-pairs occurs via a consecutive sequence

from the free, fully solvated ions (FI) to ion-pairs separated by two solvent molecules (SSIP)

and subsequently to ion-pairs with one solvent molecule separating the cation and the anion

(SIP) and finally to contact ion-pairs (CIP), in which the anions and cations are in direct

contact. The proposed sequence is shown in Figure 2-13.

Figure 2-13: Ion association mechanism as proposed by Eigen and Tamm 56, 57.

The reduced static permittivity of high-temperature water decreases the ability of water to

dissolve electrolytes. A similar trend of decreasing static permittivity with increasing

temperature has also been described for electrolyte solutions.58-60 As a result, the formation

of contact ion pairs and ionic clusters becomes more favorable at high temperatures.

Ion association in electrolyte solutions has been widely discussed in literature.61 Ion-pairs

can be studied by AC-conductance measurements, whereas with Raman spectroscopic

methods only contact ion pairs and higher clusters can be investigated, if an internal,

+ +++ - - - -

fully solvated ions ions separated by a

double solvation

layer

ions separated by a

single solvation

layer

Contact ion pair formation

Page 44: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.6 Speciation and Microscopic Structure of Salt Solutions

24

Raman-active mode is present in one or more of the constituent ions. Other spectroscopic

methods, such as neutron or X-ray scattering or X-ray absorption, can be used as well (see

Driesner62 and literature therein).

Contact ion pair formation was found from spectroscopic measurements in MgSO4

solutions.63-67 Inner- and outer-sphere complexes were studied using Raman spectroscopy

on ZnSO4 solutions to 165 °C.68 Solutions of Na2SO4 and Na2CO3 (which both exhibit type 2

phase behavior, as described above) have been studied using Raman spectroscopy, mostly at

room temperature or up to ~200 °C,69-71 but also at higher temperatures and pressures up to

>1 GPa using hydrothermal diamond anvil cells,72-75 revealing no clear evidence of contact

ion pair formation. Such fluids have also been studied by Raman spectroscopy at high

temperatures and in the presence of a controlled thermal gradient, using an optical cell

equipped with a hot rod inserted to induce precipitation.76 However, no studies have yet

investigated speciation in the liquid phase close to the three-phase (G-L-S) equilibrium line.

X-ray scattering was used to study ion association in Na2SO4 solutions at ambient

conditions,77 and X-ray absorption measurements revealed contact ion pairing in

hydrothermal halide solutions and yielded insight into the hydration structure of cations at

such conditions.78-86

The presence of larger conglomerates of ions is known from salt solutions in organic solvents

which have a low static permittivity.87 Sharygin et al.88 reported the dissociation constants

for the formation of ion-pairs in aqueous K2SO4 and Li2SO4 solutions at high temperatures

from conductance measurements. To fit the data at 400 °C they report that it is necessary to

include clusters up to Li16(SO4)80. Conductance measurements on H2SO4 solutions at high

temperatures by Hnedkovsky et al.89 also suggest the presence of clusters up to H9(SO4)5-.

Previous studies have used molecular dynamics simulations (MD) to assess water

properties90, 91 and solute speciation in hydrothermal NaCl solutions,92 phase behavior of

dense NaCl solutions at 25 MPa and up to 727 °C93 and high-temperature precipitation of

NaCl at low densities.41 In a Molecular Dynamics (MD) simulations study on the potential

mean force (PMF) of Na-Cl ion pairs, Chialvo and Simonson94 found no evidence for SSIPs at

different densities along near critical isotherms. For the SIP they found only a small

minimum in the potential, whereas the CIP seems to be the stable configuration.

Page 45: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 2

25

These studies have demonstrated that MD can be used to investigate highly dynamic

systems at extreme conditions on a molecular level, although several assumptions are

required and most models for ions and water were developed for ambient conditions. This

also leads to the results from classical MD simulations being of semi-quantitative nature.

In principle, radial distribution functions (RDFs) as obtained from MD simulations can also be

used to derive thermodynamic properties such as partial molar volumes or osmotic

coefficients, applying the Kirkwood-Buff theory.95, 96

An overview of the spectroscopic and simulation based techniques and their use to derive

thermodynamic properties of electrolyte solutions was presented by Driesner in 2013.62

Page 46: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

2.6 Speciation and Microscopic Structure of Salt Solutions

26

Page 47: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

27

Part A Macroscopic Phase Behavior of Hydrothermal

Salt Solutions

Page 48: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

28

Page 49: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

29

Studies on the Phase Behavior of Salt Solutions Using Chapter 3

High Pressure Differential Scanning Calorimetry

Parts of this chapter have been published as:

J. Reimer, F. Vogel, High Pressure Differential Scanning Calorimetry of the Hydrothermal Salt

Solutions K2SO4–Na2SO4–H2O and K2HPO4-H2O, RSC Adv., 2013, 3, 24503-24508. Used with

permission. Published by the Royal Society of Chemistry.

And:

J. Reimer, F. Vogel, Influence of Anions and Cations on the Phase Behavior of Ternary

Biomass Related Salt Solutions, J. Supercritical Fluids, 2016, 109, 141-147. Used with

permission.

3.1. Introduction

For a better understanding of the phase behavior of salt solutions we used Differential

Scanning Calorimetry as a thermal analysis tool to investigate the phase transitions in such

solutions. Thermal analysis is a sensitive and accurate tool to investigate phase transitions

and phase behavior97, 98 Differential Scanning Calorimetry (DSC) is very sensitive towards

phase transitions accompanied by changes in the heat capacity of the sample. The DSC

measures the heat flow into or out of the sample versus a reference. Changes in the heat

capacity of the sample due to phase transitions lead to step or peak shaped signals in the

heat flow curves. The measured heat flow is proportional to the sample mass m, the

heating/cooling rate and the specific heat capacity (cV) of the sample (Equation 3-1) and in

addition to that affected with an instrumental baseline.

Equation 3-1: baselinevbaselinesamplemeasured mc

As a result of this relation the signal intensity increases with high scanning (heating or

cooling) rates, but smaller thermal effects commonly vanish at such high rates due to

intense broad signals of stronger effects. This led to problems in the interpretation of the

Page 50: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.2 Methods: HP-DSC

30

heat flow curves in earlier investigations,99 in which a very high scanning rate of 10 K/min

had been chosen.

To test the method for accuracy we first measured salt solutions of Na2SO4, which have been

studied extensively before.100-110 For K2SO4 and K2HPO4 much less data is available in the

literature. Subsequently, measurements above 200°C on the ternary system Na2SO4-K2SO4-

H2O were performed.

3.2. Methods: HP-DSC

All experiments were conducted on a Sensys DSC manufactured by SETARAM, France. The

calorimeter is equipped with a Calvet-style sensor, which allows high precision

measurements, since the sensor covers almost the whole sample leading to a higher amount

of the heat flow being detected, compared to plate type DSC where the sensors only capture

the heat flow through the bottom plate of the sample container. The temperature and

energy calibration of the instrument was verified with samples of indium, tin, lead and zinc,

to cover the whole temperature range of the investigations. All metals had a purity of

> 99.999%, except indium which had a purity of > 99.995%, and were supplied by the

manufacturer of the instrument.

The salt solutions were prepared gravimetrically with an accuracy of ± 0.1 mg and then

weighed into the Incoloy crucibles with the same accuracy. Deionized water and salts with a

purity > 99.5% were used for the sample preparation. The crucibles are rated up to 50 MPa

at 600 °C. The inner volume of the crucibles was determined to be 128.63 ± 0.54 µL

(95% confidence interval) by fully filling them with water at 25 °C and weighing (see Figure

3-1). Excess water in the thread of the crucible and the safety hole at the side of the crucible

were removed by blowing with compressed air and wiping with a cellulose tissue.

The crucibles were filled with 38.6 ± 0.1 mg of the liquid sample, and therefore the average

density was 300 ± 1 kg/m3. An empty crucible was used as reference.

Page 51: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

31

Figure 3-1: Results of the measurements for the determination of the inner volume of the DSC crucibles.

During the measurement the sensor was flushed with 20 mL/min of argon. The heating rate

was 10 K/min for the heat-up phase. After an isothermal step of 1 h the measurement

started with a heating/cooling rate of 0.1 K/min.

The experimental data was treated with the Calisto Software (AKTS, Switzerland). Typically a

smoothing of 25-50 points was necessary. Multiple measurements were performed to

determine the uncertainty of the measurement.

The differences in the isochoric heat capacity (ΔcV) of the samples before and after a phase

transition were determined using the glass transition tool of the Calisto software. This tool

constructs tangents to the heat flow curves before and after a step-shaped transition and

then calculates the ΔcV from the intersection points of those tangents with a tangent on the

inflection point of the curve using the following relation:

Equation 3-2:

mc

initialfinalv

)(

Figure 3-2 shows a sketch of a typical DSC experiment with a type 2 salt solution in the

heating mode. From left to right the sequence is as follows: Initial filling at an overall density

of ~300 kg/m3. Upon raising the temperature, the density of the gas phase increases

whereas the liquid phase becomes less dense, as indicated by the darker blue for the gas

phase and the lighter blue for the liquid phase. The ions start to associate (which cannot be

determined with DSC, but with other methods, see the Part B). As the solubility limit is

127.40

127.60

127.80

128.00

128.20

128.40

128.60

128.80

129.00m

as

s (

wa

ter)

/ m

g

measurements average

95% maximum 95% minimum

Page 52: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.2 Methods: HP-DSC

32

reached a solid salt is formed and finally the gas and liquid phase coincide to a one phase

fluid. The temperature of this gas-liquid homogenization is the one, at which the last drop of

liquid evaporates into the gas phase (in case of an average density below the critical

density), or the liquid completely fills the crucible due to its thermal expansion (in case of an

average density above the critical density). If the average fluid density inside the DSC

crucible is the critical density, then the volumetric proportions between the liquid and gas

phase remain constant and the homogenization appears through a critical phenomenon, i.e.

the two phases become equal in composition and density and the phase boundary

disappears.

Figure 3-2: Sketch of a DSC experiment with a type 2 salt solution. Green and red: anions and cations; dark blue: liquid phase; light blue: gas phase, violet: single-phase fluid.

Figure 3-3: The SETARAM Sensys DSC (left) and the high pressure crucibles (right, lid, body, closed crucible, sealed and used crucible).

Runtime of the experiment

Temperature

Page 53: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

33

3.3. Results and Discussion: HP-DSC

3.3.1. Binary Solutions

3.3.1.1. Na2SO4-H2O

All the heat flow curves of the system Na2SO4-H2O show two signals. Upon heating the first

signal is an endothermic peak. This signal is attributed to the appearance of a solid phase

(G-L → G-L-S). Comparison of the heat flow curves with those obtained from experiments in

the cooling mode showed that the precipitation of the solid salt is kinetically hindered when

the measurement is performed in the heating mode and that substantial superheating

occurs. Hence, the samples were re-measured in the cooling mode, showing much lower

transition temperatures and step-like signals (Figure 3-4).

The second signal is an exothermic step and it is believed to be the disappearance of the gas

phase or the liquid phase, leading to a one phase fluid (F) in equilibrium with a solid salt

(G-L-S → F-S). The negative value of the difference in the isochoric heat capacity (ΔcV)

underlines this assumption. The temperature for this transition at an average density of

300 kg/m3 and an initial salt concentration of 1.39 mass% was 375.2 ± 0.1°C which is in

accordance with the data of Valyashko et al.,110 who obtained in a similar experiment a

transition temperature of 374.97 ± 0.17°C for a 1 mass% solution at an average density of

309.9 kg/m3. The expected λ-shape of the second signal is less pronounced in the heating

mode. In the cooling mode it is more pronounced.

Page 54: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

34

Figure 3-4: Comparison of the heat flow signals in heating and cooling mode at 0.1 K/min, 6.69 mass% Na2SO4, average density 300 kg/m3.

From the transition temperatures a part of the phase diagram can be constructed by

correcting the concentration of the liquid phase with the method proposed by Valyashko et

al.110 using the density data from Khaibullin and Novikov104 and the following equations:

Equation 3-3: MVVV vapliqvapvap

Equation 3-4:

vapliq

salt

liqliq

saltliq

VV

M

V

M

wt

x

.%100

This correction is needed because water evaporates from the starting solution during

heating and therefore the salt concentration in the liquid phase increases. Concentrations of

the salt in the gas phase are normally very small compared to the concentration in the liquid

phase and were thus neglected here.

The thermal expansion of the crucible was calculated from the material constants. The

expansion of the crucibles due to pressure changes were taken into account. Furthermore,

we relied on the experimental vapor density data available in the literature and did not use

any data on the gas density from databases or calculations. The uncertainties for the liquid

concentration were estimated by calculation of the liquid concentration for extreme values

of the liquid and vapor density, as proposed by Valyashko et al..110

320 325 330 335 340 345

T/°C

He

at

flo

w/a

.u.

Cooling

Heating

E

X

O

step-like

transition in

cooling mode

vapliq

liq

vap

MVV

Page 55: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

35

Our corrected measurements are presented in Table 3-1. A comparison with literature data

is shown in Figure 3-5. The data obtained from the DSC measurements is in excellent

accordance with the literature data.100-102, 104-108, 110, 111 The metastable liquid-liquid

immiscibility has not been observed. This is in accordance with Valyashko et al. 110

suggesting the immiscibility to be suppressed by the appearance of the solid phase.

Table 3-1: Results for Na2SO4 solutions at an average density of 300 kg/m3

initial concentration /mass%

T/°C (G-L-S) liquid solution vapor solution

x/mass % ρ/kg/m3 ρ/kg/m3

0.140 372.4 ± 0.1 0.260 ± 0.005 435 ± 2a 214 ± 2a 0.696 360.7 ± 0.2 1.02 ± 0.01 540 ± 2a 148 ± 2a 1.39 352.5 ± 0.2 1.86 ± 0.02 594 ± 2a 120 ± 2a 2.75 342.2 ± 0.1 3.47 ± 0.02 652 ± 2a 95.7 ± 2a 6.69 326.6 ± 0.3 7.96 ± 0.05 755 ± 2a 70.6 ± 2a 15.0 305.7 ± 0.1 17.1 ± 0.1 884 ± 2a 50.5 ± 2a

a Obtained from interpolation of data from Khaibullin and Novikov 104

Figure 3-5: Temperature-composition plot for the liquid at the three-phase equilibrium (G-L-S) of the system Na2SO4-H2O.

300

310

320

330

340

350

360

370

380

0 2 4 6 8 10 12 14 16 18

x/mass%

T/°

C

this study

Valyashko et al. 2000

Smits & Wuite 1909

Schoeder et al. 1935

Schoeder et al. 1937

Benrath et al. 1937

Benrath 1941

Keevil 1942

Boot et al. 1950

Marshall et al. 1954

Khaibullin & Novikov 1973

Page 56: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

36

3.3.1.2. K2SO4-H2O

K2SO4 is a type 2d’ salt like Na2SO4 and therefore the DSC experiments showed quite similar

results. Interestingly, the transition temperatures for the appearance of the solid phase are

much higher compared to the sodium salt. The liquid concentration was corrected with the

method described above. Liquid density data was extrapolated from the data of

Puchkov et al..112 Because of the lack of data for the vapor density, we extrapolated the

vapor pressures for the saturated solutions from the data of Ravich et al.113 and used the

corresponding vapor densities for pure steam obtained from the NIST database.5 The

extrapolation of the liquid densities and the vapor pressures is subject to some uncertainty,

which we assume to be of the same magnitude as the uncertainty for the density

measurements for the Na2SO4 system. Therefore we assumed an uncertainty for the

densities of ± 2 kg/m3. The uncertainty in the liquid concentration was estimated as

mentioned before. The measured and calculated values are shown in Table 3-2.

Table 3-2: Results for K2SO4 solutions at an average density of 300 kg/m3

initial concentration /mass%

T/°C (G-L-S) liquid solution vapor solution

x/mass % ρ/kg/m3 ρ/kg/m3

0.864 371.7 ± 0.2 1.69 ± 0.03 565 ±2 a 196 ± 2b 1.72 367.7 ± 0.1 2.90 ± 0.04 586 ± 2a 173 ± 2b 3.40 362.3 ± 0.6 5.19 ± 0.05 615 ± 2a 148 ± 2b 5.04 358.9 ± 0.1 7.34 ± 0.07 639 ± 2a 137 ± 2b 8.19 353.5 ± 0.1 11.4 ± 0.1 691 ± 2a 121 ± 2b

a Obtained from extrapolation of data from Puchkov et al. 112 b Obtained from extrapolation of the vapor pressures from Ravich et al. 113 and calculation by the NIST Chemistry Webbook5

Our results are in good accordance with the few data sets available in the literature 101, 113, 114

(Figure 3-6), especially with the data of Ravich and Borovaya.114 Further measurements of

densities are required to improve the correction of the measured DSC data. The DSC

measurements also showed substantial superheating in this system, therefore all data was

measured in the cooling mode.

Page 57: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

37

Figure 3-6: Temperature-composition plot for the liquid at the three-phase equilibrium (G-L-S) of the system K2SO4-H2O.

3.3.1.1. K2HPO4-H2O

In the heating mode the heat flow curves showed an endothermic step-like transition with a

positive ΔcV value followed by a second exothermic step with a negative ΔcV value (Figure

3-8). The first transition is assigned to the appearance of a second liquid phase, leading to a

liquid 1 – liquid 2 – gas equilibrium. The second transition is assigned to the critical endpoint

R, where the dilute liquid phase and the gas phase become equal. Supersaturation was also

observed for the first transition, but to a much lower extent compared to the type 2 salts.

For a solution with the concentration of 1.72 mass% the signal appeared at 372.3 ± 0.9°C in

the heating mode and at 371.4 ± 0.1°C in the cooling mode. Hence, all the temperatures

given in Table 3-3 were measured in the cooling mode.

As in the measurements with the sulfate solutions the λ-shape of the second transition was

less distinct in the heating mode compared to the cooling mode.

Table 3-3: Results for K2HPO4 solutions at an overall density of 300 kg/m3

Initial conc./mass% T/°C (G-L1-L2)

0.434 376.7 ± 0.1 0.865 374.8 ± 0.1 1.72 371.4 ± 0.1 3.96 366.6 ± 0.1 6.46 364.2 ± 0.2

Page 58: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

38

No phase density data is available in the literature for this system and therefore the

correction for the liquid phase concentration could not be applied. Comparing our

measurements to those of Marshall et al.,115, 116 which were also taken at isochoric

conditions, but probably at higher densities, we see a consistent trend of the data points

(Figure 3-7). This is especially the case if we assume that all data points in this study have to

be corrected to higher concentrations.

Figure 3-7: Temperature-initial composition plot for the liquid at the three-phase equilibrium (G-L1-L2) of the system K2HPO4-H2O.

3.3.2. Ternary Solutions

First DSC measurements on ternary salt solutions have been carried out during our first

calorimetric study,117 which focused mainly on the method validation, solubility and the

different signal types, depending on the type of phase behavior. The results obtained for the

Na-K-SO4-H2O system led to a systematic investigation of the dependence of the phase

behavior on the different ions in solutions carried out measuring different solutions at a

fixed average fluid density of 300 kg/m3, a fixed concentration of 0.1 mol/kg(H2O) and molar

ratios between the two salts of 100:1 to 1:100. We investigated salts with Na+, K+ and Mg2+

as cations and SO42-, HPO4

2- and CO32- as anions. An overview of the results is presented in

Table 3-5. During the course of our investigation we discovered that for systems containing

either MgSO4 or Na2HPO4, higher heating/cooling rates of 1 K/min (compared to 0.1 K/min)

were needed to reveal the signals of solid precipitation or liquid immiscibility. This is also

due to the signal intensity being linearly dependent from the heating rate in thermal

358

360

362

364

366

368

370

372

374

376

378

0 10 20 30 40 50 60

initial x/mass%

T/°

C

this study

Marshall 1982

Marshall 1981

Page 59: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

39

analysis. In our earlier investigation117 the measurement error for the temperature at a

heating/cooling rate of 0.1 K/min was generally around ± 0.2 °C. As a higher heating/cooling

rate most likely will induce a higher error, a maximum error for the measured phase

transition temperatures can be estimated to ± 1-2 °C. The values for the boundary binary

systems were either interpolated from our earlier work117 (Na2SO4-H2O, K2SO4-H2O, and

K2HPO4-H2O) or, in case of MgSO4–H2O and Na2HPO4-H2O, newly measured.

Table 3-4: Results of the measurements of the binary salt solutions.

Salt Solid precipitation temperature / °C

Liquid immiscibility temperature / °C

Gas – liquid homogenization temperature / °C

MgSO4 233.5* 309.5* (ms) 375.0 Na2SO4 351.9‡ - 375.1 K2SO4 367.7‡ - 376.9

Na2HPO4 - 346.4* 374.0 K2HPO4 - 371.5‡ 378.5

* measured at a heating/cooling rate of 1 K/min ‡ obtained from interpolation of earlier measurements 117 (ms) metastable

Page 60: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

40

Table 3-5: Results of the measurements of the ternary salt solutions at constant total molality of 0.1 mol (salt)/kg (water).

Salt 1 Salt 2 Molar ratio

Salt 1 : Salt 2

Solid precipitation temperature

/ °C

Liquid immiscibility temperature

/ °C

Gas – liquid homogenization temperature / °C

Na2SO4 K2SO4

100:1 352.9 - 375.1 10:1 355.8 - 376.0 1:1 368.5 376.5

1:10 374.0 369.9 377.1 1:100 368.1 - 376.6

Na2SO4 MgSO4

100:1 352.3 - 375 10:1 356.5 - 375.1 1:1 352.1 - 374.8

1:10 228.7* - 374.6 1:100 222.5* - 373.9

K2SO4 MgSO4

100:1 368.6 - 376.7 10:1 368.8 - 376.7 1:1 251.0*/374.0 - 376.6

1:10 226.1* - 374.3 1:100 213.7*/231.1* - 372.1

K2CO3 Na2CO3

100:1 - - 391.7‡

10:1 - - 389.6‡

1:1 - - 384.7‡

1:10 361.2 - 380.4

1:100 358.0 - 378.0

Na2SO4 Na2CO3

100:1 352.2 - 375.1 10:1 353.4 - 375.8 1:1 357.9 - 377.4

1:10 360.7 - 377.9 1:100 358.1 - 377.9

Na2SO4 Na2HPO4

100:1 353.4 375.2 10:1 354.2 375.3 1:1 361.6 347.3* 375.3

1:10 345.5* 374.8 1:100 332.1* 374.8

K2SO4 K2HPO4

100:1 367.0 376.7 376.1 10:1 368.9 373.0 377.6 1:1 372.4 377.5

1:10 372.5 377.6 1:100 376.7 377.7

* determined at a heating/cooling rate of 1 K/min ‡ superimposed by a liquid immiscibility

Page 61: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

41

3.3.2.1. Na2SO4-K2SO4-H2O

Interestingly, in our investigations of the ternary system K2SO4-Na2SO4-H2O we found heat

flow curves resembling those of the measurements with the K2HPO4-H2O system. The first

signal upon heating is an endothermic step with positive ΔcV value and the second signal is

an exothermic step with negative ΔcV (Figure 3-8). Therefore we assume that liquid-liquid

immiscibility occurs in mixed solutions of K2SO4 and Na2SO4. The λ-shape of second

transitions was not clearly visible in the heating mode. We did not see any precipitation of a

solid phase, but only the liquid-liquid immiscibility, in the whole concentration range of our

investigations (Table 3-6). Thus, we conclude that the ternary solutions of potassium and

sodium sulfate behave like a binary type 1d solution.

The immiscibility may arise through the metastable states which are reported for the binary

solutions of sodium and potassium sulfate. A similar effect has been reported for the system

K2SO4-Li2SO4-H2O.17, 18, 118 This liquid-liquid immiscibility was observed at all molar ratios K:Na

from 1:2 to 4:1.

Table 3-6: Results for K2SO4-Na2SO4-H2O at an overall density of 300 kg/m3, measured in the heating mode.

Initial conc. (Na2SO4) /mass%

Initial conc. (K2SO4) /mass%

Molar ratio K:Na T /°C (G-L1-L2) T /°C (G-L)

0.699 1.71 2:1 367.79 ± 0.07 376.93 ± 0.03 1.40 0.857 1:2 363.20 ± 0.06 376.25 ± 0.01 1.39 1.70 1:1 363.54 ± 0.07 376.52 ± 0.02

0.688 3.38 4:1 367.09 ± 0.01 376.99 ± 0.01

Page 62: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

42

Figure 3-8: Comparison of the heat flow curves of a K2HPO4 (6.46 mass%) solution and a Na/K SO4 (1:1) solution (1.39 mass% Na2SO4, 1.70 mass% K2SO4).

Figure 3-9: Phase transitions in the ternary system Na+-K+-SO42--H2O.

The investigation at a constant total molality of 0.1 mol (salt)/ kg (water) confirmed the

earlier findings, that the system Na-K-SO4-H2O (Figure 3-9) exhibits a stabilization of the

liquid immiscibility, which is metastable in the binary mixtures. The stabilization is present at

Na+:K+ molar ratios between 1:1 and 1:10. At a molar ratio of Na+:K+=1:1, the transition

temperature, where the second liquid phase forms is at 368.5 °C, whereas in our earlier

work117 it was found at 363.54 °C, owing to the lower total salt concentration in the present

360 362 364 366 368 370 372 374 376 378

T/°C

He

at

flo

w/a

.u.

Dipotassium phosphate

Sodium potassium sulfate (1:1)

E

X

O

350

355

360

365

370

375

380

1:0 100:1 10:1 1:1 1:10 1:100 0:1

T /

°C

Molar ratio Na2SO4:K2SO4

solid precipitation

liquid immiscibility

G=L homogenization

Page 63: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

43

study (0.1 m versus ~0.2 m, see above). In addition to the stable liquid immiscibility, a

salting-in effect can be seen in our results. The temperature at which the solid salt

precipitates increases from 351.9 °C for the binary sodium sulfate solution to 355.8 °C in the

mixture with a Na+:K+ ratio of 10:1. This salting-in effect is in accordance with the results of

Freyer and Voigt,119 who had studied the solubility in this system up to 200 °C. In mixed

solutions they reported the formation of glaserite (K3Na(SO4)2) with molar ratios of K:Na

between 1.31 and 2.91.

In the measurement with a molar ratio of Na+:K+=1:10, solid precipitation occurs at

temperatures above the liquid immiscibility. From our measurements it is not clear if the

liquid immiscibility is not fully stabilized, or if the point of the salt precipitation indicates the

presence of a G-L1-L2-S equilibrium.

3.3.2.2. Na2SO4-Na2CO3-H2O

Figure 3-10: Phase transitions in the ternary system Na+-SO42--CO3

2--H2O.

In contrast to the system K-Na-SO4-H2O, the system Na-SO4-CO3-H2O (Figure 3-10) does not

feature a stabilization of the liquid immiscibility. A slight salting-in effect can be seen for

Na2SO4, a stronger one for Na2CO3, but the solid precipitation happens prior to the G=L

homogenization in all measurements, hence all samples exhibit a pseudo-type 2 behavior,

350

355

360

365

370

375

380

1:0 100:1 10:1 1:1 1:10 1:100

T /

°C

Molar ratio Na2SO4:Na2CO3

G=L homogenization

solid precipitation

Page 64: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

44

i.e. solid precipitation at temperatures below the gas-liquid homogenization and no liquid

immiscibility. The smaller increase in the precipitation temperature in the samples with

Na2SO4 as a major compound can be attributed to the formation of a stable double salt

Na2CO3∙2 Na2SO4, as known from literature.107, 120 The data of Schroeder et al.107, 120 also

exhibit a solubility maximum at SO42-:CO3

2- molar ratios between 1:1 and 1:10.

Figure 3-11: Solubility in the Na2SO4-Na2CO3-H2O system. Data taken from Schoeder et

al..120 Note the salting-in effect on the side with Na2CO3 as major component and the

solubility minima around the tie line of the stable double salt.

0 10 20 30 40

200°C 300°

250°C

90

80

70

10

20

30

100

60 40

0 350°C

wt. % Na2CO3

Page 65: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

45

3.3.2.3. Na2SO4-MgSO4-H2O

Figure 3-12: Phase transitions in the ternary system Na+-Mg2+-SO42--H2O.

The measurements in the Na-Mg-SO4-H2O system (Figure 3-12) reveal a slight salting-in

effect for the sodium sulfate upon addition of magnesium. The solid precipitation

temperature increases from 351.9 °C for the binary Na2SO4-H2O system to 356.5 °C at a

Na2SO4:MgSO4 ratio of 10:1, hence the effect is similar to the one observed in the Na-K-SO4-

H2O system. Our measurements a high Na2SO4:MgSO4 ratios did not reveal any precipitation

of MgSO4 at low temperatures.

Interestingly, there seems to be a salting-out effect for the MgSO4 upon addition of sodium,

as the precipitation temperature decreases from 233.5 °C for the binary mixture to 222.5 °C

in the Na2SO4:MgSO4 = 1:100 mixture. This effect could be caused by the formation of a

stable mixed solid phase in the ternary mixture. The formation of mixed salts was

reported121 from solubility measurements at temperatures up to 210 °C, where the solid

phase consisted of 3 Na2SO4∙MgSO4 (vanthoffite). Note that the low temperature

precipitation temperatures were determined in the heating mode at a heating rate of

1 K/min, hence the error of those measurements being in the range of 1-2 degrees

Centigrade. The metastable liquid immiscibility in the binary MgSO4-H2O system occurs at

220

240

260

280

300

320

340

360

380

1:0 100:1 10:1 1:1 1:10 1:100 0:1

T /

°C

Molar ratio Na2SO4:MgSO4

solid precipitation

liquid immiscibility (ms)

G=L homogenization

Page 66: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

46

309.3 °C, which is in reasonable accordance with the observations of Wang et al.,64 who

visually observed the transition at 304.5 °C in a 1.19 wt.% (0.1 m) solution and characterized

the system using Raman spectroscopy.

3.3.2.4. Na2SO4-Na2HPO4-H2O

Figure 3-13: Phase transitions in the ternary system Na+-SO42--HPO4

2--H2O.

The Na-SO4-HPO4-H2O system (Figure 3-13) features solid precipitation, being present

between SO42-:HPO4

2- ratios of 100:1 and 1:1 mixtures. In the 1:1 mixture the liquid

immiscibility is present below the solid precipitation temperature. At higher HPO42- contents

the solid precipitation is absent, hence leading to a pseudo-type 1 behavior. Interestingly,

the phase separation temperature for the liquid immiscibility drops from 346.4 °C in the

binary Na2HPO4-H2O system to 332.1 °C in the ternary mixture with a molar ratio

SO42-:HPO4

2- = 1:100, but increases then to 347.3 °C in the 1:1 mixture. Note that the liquid

immiscibility in the Na2HPO4 containing solutions was only visible in the heat flow curves,

when the heating rate was increased to 1 K/min and in the heating mode only, hence also

the error in temperature for those transitions is higher than for the solid precipitation which

could be determined in the cooling mode at 0.1 K/min.

330

335

340

345

350

355

360

365

370

375

380

1:0 100:1 10:1 1:1 1:10 1:100 0:1

T /

°C

Molar ratio Na2SO4:Na2HPO4

solid precipitation

liquid immiscibility

G=L homogenization

Page 67: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

47

3.3.2.5. K2CO3-Na2CO3-H2O

Figure 3-14: Phase transitions in the ternary system K+-Na+-CO32--H2O.

The results of the measurements on the K-Na-CO3-H2O system (Figure 3-14) feature an

increased gas-liquid homogenization temperature compared to all other systems studied. In

addition, those signals differ in their shape from the signals observed in the other

measurements; most likely they are superimposed by the liquid immiscibility known from

the binary K2CO3-H2O system.19, 22 Solid precipitation is only visible at K+:Na+=1:10 and 1:100.

A direct comparison with the results of Urusova and Valyashko19 is not possible, as their

experiments were carried out at much higher pressures and higher concentrations.

355

360

365

370

375

380

385

390

395

100:1 10:1 1:1 1:10 1:100

T /

°C

Molar ratio K2CO3:Na2CO3

solid precipitation

G=L homogenization

Page 68: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.3 Results and Discussion: HP-DSC

48

3.3.2.6. K2SO4-MgSO4-H2O

Figure 3-15: Phase transitions in the ternary system K+-Mg2+-SO42--H2O.

In the K-Mg-SO4-H2O system (Figure 3-15) only a small salting-in effect for the K2SO4 occurs,

leading to an increase in the solid precipitation temperature from 367.7 °C in the binary

K2SO4-H2O system to 368.8 °C at a K2SO4:MgSO4 ratio of 10:1. At a 1:1 molar ratio of both

salts precipitation of one solid phase occurs at 251.0 °C and the second solid phase, probably

pure K2SO4, precipitates at 374.0 °C, just below the gas-liquid homogenization at 376.6 °C.

The experiment at K2SO4:MgSO4 = 1:100 revealed two signals for solid precipitation, one at

226.1 °C, i.e. close to the precipitation temperature expected for the binary MgSO4-H2O

system, and a second one at 213.7 °C, indicating the presence of a mixed solid phase, which

is thermodynamically favored at those conditions. The heat flow curve of the measurement

of the 1:1 mixture also shows two signals of precipitation, one at 251.0 °C and one at

374.0 °C, the later probably being the precipitation of pure Na2SO4.

200

220

240

260

280

300

320

340

360

380

1:0 100:1 10:1 1:1 1:10 1:100 0:1

T /

°C

Molar ratio K2SO4:MgSO4

solid precipitation

liquid immiscibility (ms)

G=L homogenization

Page 69: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

49

3.3.2.7. K2SO4-K2HPO4-H2O

Figure 3-16: Phase transitions in the ternary system K+-SO42--HPO4

2--H2O.

In the K-SO4-HPO4-H2O system (Figure 3-16) we found that the liquid immiscibility is stable

over a wide range of molar ratios. It is still present at SO42-:HPO4

2-=10:1, but here solid

precipitates at temperatures below the immiscibility. The phase separation temperature for

the immiscibility increases with decreasing hydrogen phosphate content, which also occurs

in the binary mixture.117, 122 Urusova and Valyashko24 reported similar results from their

measurements at much higher concentrations and hence also higher pressures and

temperatures.

3.3.3. Influence of Organics

The influence of organic compounds on the phase behavior was tested using one sample

consisting of 89 wt.% water, 10 wt.% 2-propanol, and 1 wt.% Na2SO4. The usage of 2-

propanol as a model substance has the advantage that it is liquid and fully miscible with

water, which facilitates filling the DSC crucibles. The signal shape of the data acquired from

this experiment basically resembled the one of the binary Na2SO4-H2O system, although the

signals were shifted significantly towards lower temperatures. The gas-liquid

homogenization was reached already at 363.3 °C, compared to the ~375 °C in the solution

365

370

375

380

1:0 100:1 10:1 1:1 1:10 1:100 0:1

T /

°C

Molar ratio K2SO4:K2HPO4

solid precipitation

liquid immiscibility

G=L homogenization

Page 70: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.4 Conclusions and Outlook

50

without 2-propanol. This effect is most likely related to the much lower critical temperature

of 2-propanol, which is at 235.15°C,123 compared to the critical temperature of pure water

(374 °C).

The solid precipitation temperature was determined to be at 338.6 °C, which is also much

lower than what is expected for a 1 wt.% Na2SO4 solution from extrapolation of our binary

data. The expected value would be at 356.8 °C, hence it can be concluded that 2-propanol

decreases the salt solubility drastically.

Peng124 recently reported an increase in the salt separation performance of a hydrothermal

gasification setup upon addition of 2-propanol to a mixed Na2SO4/K2SO4 salt solution.

3.4. Conclusions and Outlook

Isochoric, high-pressure differential scanning calorimetry proved to be a relatively simple

and fast method to investigate the phase behavior of aqueous salt solutions. It has the

drawback that liquid and vapor densities have to be known to correct the data for the liquid

concentration. Different phase transitions, such as the precipitation of a solid salt or the

appearance of a second liquid phase, can easily be distinguished in the heating mode. We

acquired new data in the low concentration range and reported the first calorimetric

measurements on liquid–liquid immiscibilities in hydrothermal salt solutions. Furthermore,

we were able to find such phenomena also in the ternary K2SO4–Na2SO4–H2O system, which

exhibits a pseudo type 1d behavior, even though it consists of two type 2 salts.

Our experiments carried out at saturated vapor and isochoric conditions revealed the phase

behavior in ternary salt solutions at low concentrations. For solutions containing Na2HPO4 or

MgSO4 the heating rate of the calorimeter needed to be increased from 0.1 to 1 K/min,

which suggests that those transitions are affected by a small change in the heat capacity of

the system, as the heat flux is proportional to the heating/cooling rate of the measurement.

For the systems containing magnesium this is possibly related to the precipitation of the

hydrated salt MgSO4∙H2O,100, 125, 126 whereas the anhydrous salts precipitate from aqueous

solutions of K2SO4 and Na2SO4 at high temperatures.113, 127 As mentioned in the introduction,

type 1(d) phase behavior is favorable for the SCWG and possibly also for the SCWO process.

The results indicate an order for the cations in ternary solutions for favoring type 1d

Page 71: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 3

51

behavior. The ability to form a type 1d behavior decreases in the sequence K+ > Na+ > Mg2+.

For the anions the order HPO42- > SO4

2- > CO32- can be derived. This corresponds also to the

trends for the binary solutions. For instance the solid precipitation temperature for the

sulfate salts of the above mentioned cations increases in the order Mg2+ < Na+ < K+.

Furthermore, the precipitation temperature for sodium sulfate is higher than the one for

sodium carbonate, whereas sodium hydrogen phosphate is a type 1d salt, showing no

precipitation of solid salt.

A possible explanation for this behavior is, as suggested for the binary mixtures by

Baierlein,128 the difference in the radius of the ions. Generally, the ratio of the charge to the

radius can be considered as well. A smaller ionic radius (and in case of Mg2+ also the higher

charge) leads to a stronger electrostatic interaction between the ions, which causes ion

association and clustering, prior to precipitation at high temperatures. An exception seems

to be the carbonate anion, as it has a smaller ionic radius than the sulfate anion, but in

contrast to the later one exhibits a type 1d phase behavior, if potassium is the cation in

solution. Our investigations with Molecular Dynamics simulations on sulfate solutions

showed that the association/clustering behavior is relevant for the phase behavior of type

2.129 This clustering tendency is mainly driven by the strength of the anion-cation interaction.

This leads to the hypothesis that in polyatomic anions also the number of coordination sites

plays a role, as the sulfate anion offers more possibilities for bidentate coordination of a

cation (6 in total, plus 4 for tridentate coordination, from its tetrahedral geometry),

compared to the carbonate ion (3 sites for bidentate coordination, from its planar

geometry).

Changes in the phase behavior of mixed salt solutions offer an opportunity for optimizing the

salt separation process for SCWO and SCWG applications, as type 1 salt solutions can be

separated and recovered efficiently. Hence, the addition of additional salts to a feed solution

might change the phase behavior e.g. of a type 2 solution in such a way that no solid salts

precipitate and that the separation can be conducted successfully.

It was suggested earlier130 to use phosphate salts, especially with sodium and ammonium as

cations, at concentrations of above 10 wt.% of the total salt concentration, to avoid plugging

by salt precipitation in the SCWO process. Our results reveal that the phosphate salt should

Page 72: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

3.4 Conclusions and Outlook

52

always be the major salt on a molar basis. Furthermore, based on our findings, especially the

usage of HPO42- with potassium as cations could influence the phase behavior in a favorable

way to enhance the salt separation and recovery in SCWO and SCWG processes. In

particular, the liquid immiscibility offers the possibility to efficiently recover salts from the

process stream in the form of concentrated brine.

Notwithstanding the present advances, more studies on higher order mixtures are necessary

to explore the possibilities and limits of this proposed technique of enhancing salt separation

in SCW processes. The lack of density data for gas and liquid phases at hydrothermal

conditions is currently a limitation, as this data is needed to fully interpret the isochoric DSC

measurements.

Page 73: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 4

53

Studies on the Phase Behavior of Salt Solutions with the Chapter 4

Optical Cell

4.1. Introduction

As a complimentary method to the DSC-measurements, we used visual observation, to study

the phase behavior of some of the salt solutions of interest. The confirmation of the liquid

immiscibility found in the ternary Na-K-SO4-H2O solutions was the major goal of this study.

Studying the phase behavior of salt solutions using an optical cell offers the possibility to use

either isobaric or isothermal conditions, in contrast to the HP-DSC studies presented in

Chapter 3, which are limited to isochoric measurements. Most of the experimental work in

this part has been performed by Martin Gorman in the context of an IAESTE summer

internship (June-August 2013).131

4.2. Methods: Optical Cell

For our investigations with an optical cell, we used a setup similar to the one described by

Rouff et al.99 Several modifications to the original setup have been made, e.g., the preheater

has been removed.

The pump used to fill and pressurize the cell was a syringe pump (Model 260D, Teledyne

ISCO, USA). The cell (SITEC, Switzerland) is equipped with a fluid inlet, a fluid outlet, two

sapphire windows, and a thermocouple situated in the inner cell volume (~2mL) to assure a

correct temperature measurement. This thermocouple is also connected to a heater control,

which drives four electric heating cartridges situated in the cell body. The pressure is

maintained using a manually operated backpressure valve. The setup is equipped with a

digital and an analogue pressure gauge. Pictures of the phase changes where taken via a

mirror using a LX 7 camera (Panasonic). A schematic sketch of the setup and a photograph of

the cell are presented in Figure 4-1 and Figure 4-2.

In contrast to earlier experiments with this cell by Rouff et al.,99 the experiments were

conducted in static mode, meaning that no flow through the cell was applied. This was

Page 74: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

4.2 Methods: Optical Cell

54

established through a 3-way valve, which allows switching from flow-through mode to fill

the cell with the desired solution to static mode. Measurements in the static mode avoid

possible enrichment of salt inside the cell in case precipitation conditions are being

approached. Temperature and pressure were read out to a computer.

The experiments were conducted in isothermal mode, meaning that after several minutes of

equilibration at a constant temperature, the pressure was slowly lowered manually with the

back-pressure valve until heterogenization of the fluid inside the cell was observed. The

pressure was noted manually from the display of the computer. As some types of the

heterogenization of fluids can be kinetically hindered, the pressure was raised again and also

the pressure of the homogenization was noted.

Figure 4-1: Schematic outline of the experimental setup. TC: thermocouple; PI: manometer; PTR: digital pressure gauge.

high-pressure

pump

salt solution

effluent tank

3 way valve

TC

PI

PTR

Cooler and

backpressure

regulator

flow-through mode

static mode

cell

Page 75: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 4

55

Figure 4-2: Photograph of the cell; insulation removed.

4.3. Results and Discussion: Optical Cell

4.3.1. K2SO4-H2O

To assess the performance of the optical cell, several reconnaissance experiments were

performed with K2SO4-H2O solutions at different salinities. The results reveal that the vapor

pressures of such solutions are close to those of pure water (see Figure 4-3), indicating that

the gas-liquid equilibrium region in such solutions is relatively small. Furthermore, the

results indicate that the precision of the pressure determination is prone to a relatively high

statistical error (± 6 bar) due to the manual operation of the cell, but an overall good

agreement with the NIST data. The solubility data obtained from the experiments show a

high deviation from literature data, especially at low salinities (see Figure 4-4), where the

solid precipitation should occur closer to the critical endpoint q, which is approximately

around 375 °C.

Heater connections

Fluid in- and outlets

Optical access

Page 76: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

4.3 Results and Discussion: Optical Cell

56

Figure 4-3: Comparison of experimentally determined vapor pressures with NIST data.5 Note the pressure deviations of several bar between experimental and literature data, as well as the vapor pressures of the 10 wt.% K2SO4 solution being close to the one of pure water.

Figure 4-4: Experimentally determined solubility of K2SO4.

80

100

120

140

160

180

200

220

240

290 310 330 350 370 390

p /

bar

T / °C

NIST Pure WaterData

10 % wt. K2SO4

Experimental WaterData

330

335

340

345

350

355

360

365

370

375

0 5 10 15 20 25 30

T /

°C

% wt. K2SO4

Optical Cell

HP-DSC

Ataev et al1994

Ravich &Borovaya 1968

Benrath et al1937

Page 77: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 4

57

4.3.2. NaxK2-xSO4-H2O

Studies on the liquid immiscibility in binary K2HPO4-H2O solutions were performed earlier in

our group.99 From those experiments it is known that phase separation of such fluids is not

clear in the cell used for the investigations. Rather, the liquid in the cell turns from

transparent to opaque.

Upon lowering the pressure at elevated temperature, firstly the single-phase fluid splits into

two phases, comprising a dilute liquid and a concentrated liquid (sometimes called ‘heavy

liquid’). This leads to an opaque content in the cell, as the two liquids do not separate into

two clear, translucent phases, but rather resemble an emulsion. Further lowering the

pressure leads to formation of an additional gas phase, as the G-L1-L2 equilibrium is

established. At pressures below this equilibrium the liquid becomes clear again, as the two

phase G-L equilibrium is reached. This is typical for a phase behavior of (pseudo-) type 1d.

Figure 4-5 shows the observed three phase G-L1-L2 equilibrium in an eqimolar Na2SO4-

K2SO4-H2O solution containing 0.1 mol/L of each salt, as prepared at room temperature. In

Figure 4-6 the resulting p-T diagram of this solution is presented. The lowest temperature at

which the liquid immiscibility happens, is at around 350 °C at this salt ratio and

concentration. The slope of the L1-L2 equilibrium line over temperature is rather steep,

whereas the three-phase line G-L1-L2 exhibits a relatively mild slope.

Figure 4-5: Phase transition L1-L2 to G-L1-L2 as observed in the optical cell at 212 bar and 365°C in a solution containing 0.1 mol/L Na2SO4 and 0.1 mol/L K2SO4 as prepared at 25°C. The two liquid phases are not clearly separated and resemble an opaque emulsion.

Page 78: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

4.3 Results and Discussion: Optical Cell

58

Figure 4-6: p-T phase diagram between 350 and 380 °C for a K2SO4-Na2SO4-H2O solution with a concentration of 0.1 mol/L of each salt, as prepared at 25°C. Arrows indicating the phase transitions occurring upon lowering the system pressure.

Further investigations were carried out at a constant weight fraction of 1 wt.% salt and

different Na:K ratios. The results of those experiments are presented in Figure 4-7. The data

reveal that a mixture containing 0.95 wt.% Na2SO4 and 0.05 wt.% K2SO4 exhibits a liquid

immiscibility and hence pseudo-type 1d behavior. This is especially interesting, because in

the salt separation process type 1d mixtures could be easily separated and recovered,

whereas type 2 salts (such as Na2SO4 and K2SO4 in their binary mixtures with water) lead to

accumulation of salt in the separator and hence to blocking of the separator vessel.

Figure 4-7: Results for mixed K2SO4-Na2SO4 solutions at a total concentration of 1 wt.% salt.

160180200220240260280300320340

345 355 365 375 385

p /

bar

T / °C

F → G-L

F → L1-L2

L1-L2 → G-L

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

Na 2

SO4 w

t.%

K2SO4 wt. % L-L Not Observed L-L Observed

Page 79: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 4

59

4.4. Conclusions and Outlook

Using the optical cell as a complementary method to the DSC-experiments, we investigated

the phase transitions in aqueous K2SO4 solutions. Due to the lack of an accurate pressure

determination, and due to the absence of a stirring device, which would facilitate

equilibration, the data obtained is prone to an error in the magnitude of several bar

pressure. The solubility data for concentrated solutions followed the same trend as the data

obtained from the DSC experiments, i.e. increasing solubility with decreasing salt

concentration, but the observed precipitation temperatures were approximately 4-7 °C

lower than the ones determined by DSC. This might be caused by the cell not being fully

isothermal, and the salt starting precipitation at the hottest parts of the cell. The

experiments revealed the nature of the phase transitions in ternary solutions of the system

Na2SO4-K2SO4-H2O, namely a liquid immiscibility of the d type. Hence there is evidence that

the metastable liquid immiscibility of the two subsystems is stabilized in the ternary system.

This behavior is similar to the results reported on other systems such as K2SO4-Li2SO4-H2O.17,

18, 118

Summing up on the abilities of the setup, the strength of the setup is definitively in the

observation of the nature of phase transitions. With a stirring device and an improved

temperature and pressure measurement also quantitative data could be obtained. Here also

the method of isobaric measurements is suggested using slow heating ramps, in order to be

able to determine the p-T conditions of phase transitions more accurately.

Page 80: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

4.4 Conclusions and Outlook

60

Page 81: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

61

Part B Microscopic Structure of Hydrothermal Salt

Solutions

Page 82: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

62

Page 83: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

63

Studies on the Speciation in Hydrothermal Salt Solutions Chapter 5

with X-ray Absorption Spectroscopy

5.1. Introduction

Extended X-ray absorption fine structure (EXAFS) measurements offer a unique possibility to

study the chemical surrounding of atoms in molecules and in solution. The technique has

previously been used to study salt solutions at high pressures and temperatures (see

Driesner62 and literature therein). Heavy elements are easily accessible with EXAFS

measurements, due to the fact that thick, commercially available sapphire windows and

conventional high pressure cells can be used, because of the high penetration depth of high

energy X-rays used for exciting the core electrons of those elements. For EXAFS

measurements on light elements (such as sulfur, phosphorus, potassium and sodium) special

cell designs with ultrathin X-ray transparent windows and very short pathlengths in the

sample are necessary. This causes problems in the cell construction, as those windows have

to withstand the extreme conditions of high pressures and temperatures as well as possible

corrosion from the salt solution under investigation. So far only Fulton et al.132 were

successful in measuring EXAFS-spectra at the Ca K-edge (4038.5 eV) at hydrothermal

conditions (up to 500°C, 1 kbar), using thin diamond (25 μm) windows and pathlengths of

150 μm. They claim to have a transmission of 15% at the Cl K-edge (2822 eV).

The K-edges of the elements of interest in this thesis (sulfur and phosphorus) are at even

lower energies, hence a new cell/window design was developed and tested. Unfortunately,

the new windows design was prone to leakage which prevented the acquisition of EXAFS

spectra. The approach, improvements and suggestions for future experiments will be

discussed in this part of the present thesis.

Page 84: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.2 Setup and operation description

64

5.2. Setup and operation description

The cell used in our group is similar to the one used by Fulton et al.132 An isometric

perspective of the cell and a cross-section are shown in Figure 5-1 and Figure 5-2.

The cell body is manufactured from Titanium Grade 5 and is equipped with two Swagelok

connectors for supplying the cell with feed solution. Attached to the cell body are two

heating wires. A thermocouple is inserted in a hole in the principle axis of the cell, reaching

down close to the actual cell volume.

The two windows of the cell consist of a holder unit. The diamond windows are soldered to

the tip of the titanium holders (described in detail below). Nuts, consisting of silver plated

Titanium Grade 5, are used to tighten the window holders.

The main difference compared to the cell of Fulton et al.132 is that our cell uses a distance

ring, which is in contact with both window holders, to adjust the distance between the two

diamond windows. This allows a precise adjustment of the spectroscopic pathlength.

To flush and pressurize the cell, we use a Teledyne ISCO syringe pump (Model 260D) and a

manually operated backpressure valve. At low flowrates (typically 0.1 mL/min was used), this

allowed a stable operation with pressure fluctuations below ±0.5 bar. The heating wires are

connected to a power controlled heater. During the first online tests at the beamline it

turned out, that operating the heaters at constant electrical power is preferable, as small

oscillations in the heating power, when heating to a set-temperature caused movement of

the cell due to thermal expansion of the cell and the underlying stage. With operation at

constant power this problem could be solved.

The cell was mounted onto a water cooled stage; insulation of the cell is not needed as the

cell is operated in the vacuum measurement chamber (which is needed to avoid loss of the

X-ray beam intensity due to absorption in the air), hence convectional heat loss is negligible.

A photograph of the mounted cell is presented in Figure 5-3.

Page 85: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

65

Figure 5-1: Isometric view of the Diamond Cell. 1: Cell body; 2: Insulating plate; 4: Nut; 10 and 11: Swagelok connectors.

Figure 5-2: Horizontal cross-cut through the Diamond Cell. 1: Cell body; 3: Window holder; 4: Nut; 5: Distance ring; 6: Seal; 7 and 11: Connections to Swagelok; A: Diamond windows.

Page 86: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.2 Setup and operation description

66

Figure 5-3: Diamond window cell setup as mounted in the measurement chamber of the PHOENIX beamline at Swiss Light Source.

Page 87: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

67

5.3. Improvements on Cell and Windows

5.3.1. Windows

In collaboration with Diamond Materials, Germany, newly designed diamond windows were

developed. Fulton et al.132 used diamond membranes glued to the window holders as

windows. Their membranes were 100 μm thick, which is not applicable in our case. To obtain

a reasonably high transmission for the measurements at the S K-edge the membranes

should have a thickness of less than 10 μm. Earlier tests in our group showed that the

approach of gluing such thin membranes to the window holders is not applicable. Therefore

it was decided to use a vacuum soldering technique to assemble the new window units.

The diamond membrane of ~8 μm thickness is soldered to a 0.5 mm thick diamond ring with

an outer diameter of 1.79 mm. The inner diameter of the diamond ring, i.e. the

spectroscopic aperture is 100 μm.

This unit is then soldered into a window holder made from Titanium Grade 5, which

comprises some step like shape for taking up the distance ring and also the gasket towards

the cell.

An overview of the window assembly is given in Figure 5-4. A close-up sketch of the diamond

window is presented in Figure 5-5.

Page 88: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.3 Improvements on Cell and Windows

68

Figure 5-4: Overview of the window assembly used in the Diamond Cell. Window holders in grey, distance ring in black, diamond window in blue; blue arrows: flow of liquid through the assembly; red arrows: X-ray path. Drawn to scale.

During the test of the new window design, which was performed at the PHOENIX beamline,

Swiss Light Source, Switzerland, it turned out that leakage occurred basically every time the

cell was pressurized, even though the window units were leak tested with helium

beforehand by the manufacturer. This leakage was not detectable in offline tests, as the

amount of liquid leaking out of the cell was very small and the tests were carried out with

de-ionized water, hence no fluid drops and/or salt precipitates could be visually detected.

During operation of the cell at the PHOENIX beamline using salt solutions, this small amount

of liquid leaking through the windows was enough to lead to salt deposition in the pinhole of

the diamond ring of the window and also in the upper part of the window holder. The

precipitated salt fully blocked the optical path and made spectroscopic measurements

impossible.

Page 89: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

69

Figure 5-5: Close-up view of the window assembly as used in the Diamond Cell. Window holder in grey, diamond ring and membrane in blue, solder in orange. Drawn to scale.

Examination of the windows under the microscope showed that the unused windows exhibit

a clean intact surface of the diamond membrane (Figure 5-6). The used windows, which

exhibited leaking, showed cracks in the diamond membrane in the vicinity of the pinhole of

the diamond ring (Figure 5-7).

Analysis using the X-ray fluorescence of sulfur (Figure 5-8), after a test with a 10 wt.% Na2SO4

solution, revealed that even after extensive rinsing of the window with de-ionized water,

sulfur can be detected on and underneath the diamond membrane, which suggests that the

sulfate migrated through the cracks in the membrane, probably also leading to corrosion of

the solder material.

1mm

Page 90: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.3 Improvements on Cell and Windows

70

Figure 5-6: Microscopy of an unused diamond window, showing the clean and intact diamond membrane, soldered onto the diamond ring (outer diameter 1.79 mm, inner diameter/optical aperture 0.1 mm).

Figure 5-7: Microscopy of a broken diamond window. Note the cracks in the diamond membrane close to the center, where the pinhole is underneath.

Diamond ring

Diamond membrane

Optical aperture

Page 91: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

71

Figure 5-8: Fluorescence map of a broken diamond window after intensive rinsing with de-ionized water. Sulfur in red, silver (solder material) in blue.

5.3.2. Distance Rings

The vertical and horizontal alignment of the window units is obtained through a distance

ring, which has two 1 mm holes at the sides through which a fluid can be introduced

between the two diamond windows.

New distance rings were manufactured by Diamond Materials from Titanium Grade 5. To

ensure the desired spacing between the two diamond membranes of 10-50 μm, the distance

rings and the window units were gauged by messtec gmbh, Switzerland. The thickness of the

distance ring, together with the vertical distance between the diamond membrane and the

first step of the window holder allowed the calculation of the inter-window distance. The

results of the measurements and the calculated distances are given in Table 5-1 and Table

5-2.

Table 5-1: Measured vertical distances between the diamond membranes and the first step of the holder for the four window units manufactured by Diamond Materials. Values given in mm.

Window unit no.

I II III IV

min 3.090 3.094 3.157 3.2193 max 3.092 3.096 3.160 3.2233

Optical aperture

Page 92: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.3 Improvements on Cell and Windows

72

Figure 5-9: Detail of the window unit with indicated gauged distance between window and the first step of the holder (dWindow).

Table 5-2: Calculated distances between the diamond windows. dtotal is the sum of the measured distances between the windows and the first step of the holder (dWindow) for both windows of the respective combination. Green background: suitable combinations for conducting experiments; Yellow background: Distances too large; Red background: Negative distances, danger of damaging the windows.

window unit

Thickness of distance ring /mm

combination dtotal /mm 6.299 6.238 6.429 6.244 6.408

I-II 6.188 0.111 0.050 0.241 0.056 0.220 I-III 6.252 0.047 -0.014 0.177 -0.008 0.156 I-IV 6.3153 -0.016 -0.077 0.114 -0.071 0.093 II-III 6.256 0.043 -0.018 0.173 -0.012 0.152 II-IV 6.3193 -0.020 -0.081 0.110 -0.075 0.089 III-IV 6.3833 -0.084 -0.145 0.046 -0.139 0.025

5.3.3. Gaskets

Previous tests with silver gasket rings revealed that at high temperatures, especially above

the critical point of pure water (>374 °C), leakage occurred. To overcome this, we used

copper rings with rectangular cross-section. The copper rings were heated to bright red heat

and shock-cooled via immersion into ethanol (96% v/v). This procedure yielded soft copper

with a shining surface.

dWindow

Page 93: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 5

73

For assembling the cell, the gasket ring was inserted first on one side of the cell. Then the

first window unit was inserted and the nut was hand-screwed into the cell. Upon rotation of

the cell, the distance ring was inserted, using tweezers. Afterwards followed the second

gasket, the second window unit and the second hand-tightened nut. The nuts were

alternately tightened using a torque wrench with increasing torque (maximum torque

applied was 120 Nm), to assure the same pressure on both sealing rings.

The whole procedure led to reliable tightness of the cell up to ~380 to 390°C at 300 bar. At

higher temperatures leakage still occurred.

5.4. Conclusion and Recommendations

The measuring of the thickness of the distance rings and the vertical distance between the

diamond membrane and the first step of the window holder proved to be a reliable process

to assure a defined distance between the windows, i.e. the spectroscopic pathlength.

A revised design of the cell body should comprise a better design for the sealing. The use of

metal V- or C-rings, possibly spring loaded and/or with internal pressure loading is favorable.

Those seals are also used in industrial applications for similar pressure and temperature

conditions and hence are commercially available at moderate cost.

The new design of the diamond windows was promising, as it overcomes the unreliable

gluing of diamond membranes to the window holder. But due to constant leakage problems,

it is not clear if this approach will be successful eventually. One reason might be the

extremely thin diamond membrane, which is probably not capable of withstanding the

pressures applied in our measurements (up to 300 bar).

One approach to overcome this would be using X-ray Raman scattering, which allows the

usage of higher energy X-rays, hence no ultra-thin windows have to be used and the cell

design can be conventional, using known methods and gaskets. This method has already

been applied e.g. to examine the structure of supercritical water,133 but needs a special X-ray

Raman detector, which is currently not available at the PHOENIX beamline.

Page 94: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

5.4 Conclusion and Recommendations

74

Page 95: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

75

Studies on the Speciation in Hydrothermal Salt Solutions Chapter 6

with FTIR-Raman Spectroscopy

Parts of this chapter have been published as:

J. Reimer, M. Steele-MacInnis, J.M. Wambach, F. Vogel, Ion Association in Hydrothermal

Sulfate Solutions, studied by Modulated FT-IR-Raman and Molecular Dynamics, J. Phys.

Chem. B, 2015, 19 (30), 9847–9857. Used with permission. Copyright 2015 American

Chemical Society.

6.1. Introduction

Raman spectroscopy is a sensitive tool to investigate speciation in electrolyte solutions,

provided that at least one of the ions has internal vibrational modes that are altered when

ion pairs and clusters are formed. For our investigations, the sulfate salts are most

interesting, as they exhibit a phase behavior of type 2d’. We studied the speciation close to

precipitation conditions at high temperatures in order to reveal species relevant for the

decreasing solubility at high temperatures. In contrast to the XAFS-measurements, Raman-

spectroscopy does not yield any direct structural information on the species in solution,

which makes the use of additional methods necessary (See next part of this thesis).

So far Raman spectroscopy has been used to study sulfate solutions either at very high

pressures and temperatures,72 at high concentration at ambient conditions67, 69 or at

conditions, where precipitation already takes place.76

In this part of the thesis, we present our measurements on different type 2d’ salt solutions

close to precipitation conditions with the main focus on Na2SO4-H2O and in addition also

measurements on Li2SO4-H2O, Na2CO3-H2O, and the type 1d system Na2HPO4-H2O close to

conditions, where the liquid immiscibility occurs.

Page 96: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.2 Methods: FTIR-Raman Spectroscopy

76

6.2. Methods: FTIR-Raman Spectroscopy

We use a Thermo Fisher Nicolet 8700 FT-IR spectrometer in combination with a NXR Raman

unit with a 2.5 W Nd:YVO4 laser (1064 nm) for spectroscopic measurements (Figure 6-2). An

optical chopper is installed in the sample compartment, as well the high-pressure, high-

temperature cell. The chopper is connected to a control unit and the chopper frequency is

fed to a lock-in amplifier (SR 530, Stanford Research, USA). The backscattered (180°

geometry) Raman signal passes through the Michelson interferometer of the Nicolet 8700

equipped with a CaF2 beamsplitter, and is recorded by a InGaAs detector. The detector signal

is fed to the lock-in amplifier, using the output channel of the Nicolet spectrometer, and is

transferred back after the demodulation. The measurements are carried out using the Step-

Scan mode of the spectrometer with a resolution of 8 cm-1. The averaging time for each

interferogram point is typically 2000 ms after a settling time of the mirror of 100 ms. 4640

points are measured per interferogram, which takes ~2:45 h. The chopping frequency is set

to 2000 Hz and the filter of the lock-in amplifier is set to 1 ms. The measured spectra are

analyzed using the OMNIC software supplied by Thermo Fisher and applying Happ-Genzel

apodization for the Fourier-transformation. The deconvolution of the spectra was done

without smoothing, applying a linear baseline and using Voigt-functions for the peak fitting.

Page 97: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

77

Figure 6-1: Schematic sketch of the signal generation and processing in the modulated FTIR-Raman spectroscopy.

Figure 6-2: Overview of the experimental setup for FTIR-Raman spectroscopy.

Nd:YVO4-Laser (1064nm)

Chopper Sample Edge filter

SST-Michelson-Interferometer

InGaAs-Detector

Lock-in amplifier

Data recording (Interferogram)

Fourier-Transform

Final Spectrum

νm

Raman signal is modulated with νm. Thermal background stays unmodulated

Periphery see Fig. 6-3.

Page 98: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.2 Methods: FTIR-Raman Spectroscopy

78

Figure 6-3: Periphery of the FTIR-Raman setup.

The high-pressure cell with an internal volume of 0.8 mL is supplied by SITEC AG, Switzerland

and equipped with a 8 mm thick sapphire window with a free aperture of 8.5 mm. The cell

can be operated up to 400°C and 300 bar (see Figure 6-4). After flushing the cell with the salt

solution, the pressure is kept constant by closing the back pressure valve and using the

constant pressure mode of the high pressure syringe pump (D260, Teledyne ISCO, USA). The

heating cartridges inside the cell walls are controlled via a PID controller, connected to a

computer, which also records the pressure and the internal temperature of the cell (Figure

6-3). Temperature is measured using a calibrated thermocouple placed in the middle of the

cell. Using this configuration the pressure and temperature are stable within ± 0.1 bar and ±

0.1°C, respectively, whereas the absolute errors for pressure and temperature are ± 0.5 bar

and ± 0.3°C.

The salts were supplied by various suppliers and had a purity of >99.5%. Solutions were

prepared gravimetrically with de-ionized water.

Page 99: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

79

Figure 6-4: Photograph of the high-pressure cell mounted in the experiment chamber of the NXR Raman module (left) and technical drawing of the high pressure cell (right). Red lines represent the optical path of the infrared light.

Page 100: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.3 Results and Discussion: FTIR-Raman Spectroscopy

80

6.3. Results and Discussion: FTIR-Raman Spectroscopy

6.3.1. Na2SO4-H2O

Experiments were conducted using a 10 wt.% solution. For this composition, the

precipitation temperature at vapor pressure is 320.3 °C, according to interpolation from data

available in literature.102, 104-106, 110, 111, 117 Spectra were collected at 310, 312, 314, 316 and

318 °C. Pressure was set to 12 MPa. This pressure is approximately 1 MPa above the vapor

pressure of pure water at the temperatures of our experiments.5

The total symmetric stretching mode ν1 of sulfate occurs at 981.5 cm-1 at 20°C and

atmospheric pressure, but shifts to 974 cm-1 at 300 °C and 9 MPa, and at higher

temperatures to 973 cm-1 and 971 cm-1 in the spectra collected at 316 and 318 °C,

respectively. For simplicity, this band is referred to as the ca. 973 cm-1 band hereafter. The

assignment of this Raman band to the ν1 vibration of sulfate is supported by previous

studies.63, 72, 76 Unfortunately, this Raman feature does not allow one to distinguish

contributions from truly unassociated ions (free ions, FI) versus solvent-shared (SIP) and

solvent-separated ion pairs (SSIP).72 Consequently, we cannot quantify the proportions of FI,

SIP and SSIP in the present study, and it is likely that all of these species contribute to the

observed Raman band.

The shift of the ν1-band from 981.5 cm-1 at room temperature to 971 cm-1 at 318 °C could be

explained either by a change in the S-O stretching frequency owing to the higher

temperature, or an increasing proportion of small clusters (corresponding to the MD

simulations at temperatures below 280 °C; see section Chapter 7). Zhang et al.134 calculated

a small shift to lower frequencies in the sulfate stretching mode for the NaSO4– ion pair and

the Na(SO4)23- triple ion using density functional theory, supporting the latter explanation.

Nevertheless it is likely that both effects (ion association and temperature change)

contribute to the observed shift. Watanabe and Hamaguchi135 showed that a shift in the

sulfate Raman band can also be attributed to dynamics of ion association in aqueous MgSO4

solutions on the pico- to femto-second timescale.

Page 101: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

81

With increasing temperature, a new Raman band appears at ca. 993 cm-1 and increases in

intensity (Figure 6-5). We consider that this Raman band represents the ν1 vibration of

sulfate in ionic clusters. This vibrational band is close to that of crystalline Na2SO4 phase I

(993.5 cm-1), the stable solid phase of Na2SO4 at the conditions of our experiments.136 Note,

however, that the experimental temperature-pressure-composition conditions of our

experiments are such that precipitation should not occur (i.e., the solution is undersaturated

with respect to Na2SO4) and other signals of the solid phase are not present in the spectra. In

addition, the increase of the signal at 993 cm-1 appears over a temperature range of several

degrees Celsius, which would not be expected if this reflected the precipitation of a solid

salt. Furthermore, reconnaissance experiments at equivalent pressures and temperatures as

presented in Figure 2, but conducted using a flow-through setup, showed no evidence of salt

precipitation (which would have been easy to detect, owing to buildup of salt during flow-

through, or pressure fluctuations resulting from clogging). Therefore we are confident that

salt precipitation did not occur in the present experiments, and that the 993 cm-1 band

represents aqueous species favored at elevated temperature. Our assignment of the

ca. 993 cm-1 band to small and medium-sized clusters is supported by studies of MgSO4-

bearing systems, in which the signal of contact ion pairs and higher aggregated clusters

appears on the higher frequency side compared to the signal of the sulfate ion.63, 64, 69

Furthermore, theoretical calculations by Zhang et al.134 predict a shift of the Raman band of

sulfate towards higher wavenumbers in the triple ion Na2SO40 and the cluster Na3SO4

+.

Remarkably, the predicted shift134 in the Raman peak position with the transition to triple

ions plus clusters is ~30 cm-1, close to the measured shift of ~20 cm-1 in the present study.

Note that Wang et al.64 observed increased ion association and polymerization in the

MgSO4-rich liquid formed via (possibly metastable) liquid immiscibility in the H2O-MgSO4

system; in the present study of Na2SO4 solutions we can exclude such liquid immiscibility,

based on earlier work117 that showed that liquid-liquid phase separation in Na2SO4 solutions

occurs only in ternary (or higher) systems, and at temperatures >360 °C.

Page 102: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.3 Results and Discussion: FTIR-Raman Spectroscopy

82

Figure 6-5: Raman spectra showing the ν1 stretching mode of the sulfate anion at a pressure of 120 bar and temperatures of 308, 310, 312, 314, 316 and 318°C.

Matsumoto et al. presented a Raman spectrum (Fig. 6b in ref 76) of a Na2SO4 solution

measured close to the heated rod, upon which solid Na2SO4 precipitated, in an optical cell.

The main signal in the Raman spectrum76 of the solution is that of SO42- at ~973 cm-1, but a

shoulder around 993 cm1 is also visible in the spectrum, close to the position of the solid

Na2SO4 ν1 vibration (similar to our spectrum taken at 314 °C, 12 MPa; see Figure 6-5). Based

on the results of the present experiments (in which the ca. 993 cm-1 band appears prior to

solid precipitation), we suggest that the shoulder at ca. 993 cm-1 in the solution spectra from

experiments by Matsumoto et al.76 represented ion association near the liquid-solid

interface, similar to that observed prior to onset of precipitation in the present study.

As a measure of the proportion of ionic clusters in solution, we calculated relative integrated

peak areas of the two (ca. 973 and 993 cm-1) Raman bands. Note that the relative peak areas

do not represent the relative concentrations of associated versus non-associated species,

owing to the different molar Raman scattering efficiencies (i.e., Raman cross sections) of the

different clusters. However, the variation in relative peak areas likely reflects trends (at least

qualitatively) in the proportions of associated versus unassociated species. The

Page 103: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

83

deconvolution of the signals showed that the relative peak area of the clusters increases

with increasing temperature from 0% at 310 °C to 67% at 318 °C (Figure 6-6).

Figure 6-6: Temperature dependence of the relative peak areas of the ca. 973 and 993 cm-1 Raman peaks in a 10 wt.% Na2SO4 solution at 120 bar. The variation in relative peak areas qualitatively reflects the relative abundances of free (unassociated) ions and ion pairs versus small and medium sized clusters.

The influence of pressure on cluster formation was also studied, by measuring Raman

spectra at 316 °C and 11, 12, 13, 15 and 22 MPa (Figure 6-7). Deconvolution of the signals

indicates a decrease in the relative peak area representing clusters from 57% at 11 MPa to

10% at 15 MPa and further to 0% at 18 MPa (Figure 6-8).

An earlier spectroscopic study of sodium sulfate solutions in a hydrothermal diamond-anvil

cell72 did not find any signal comparable to the ca. 993 cm-1 band reported here. This result

suggests that contact ion pair formation was absent in sodium sulfate solutions in the

previous study.72 We suggest that the difference in ion association in these studies likely

reflects differences in pressure. Specifically, the study by Schmidt72 was conducted at

pressures significantly greater than in the present study (e.g., 165 MPa at 300°C,72 versus

<22 MPa in the present study). Thus, the solvent dielectric constant was concomitantly

higher in the former study,72 limiting the extent of ion association. Also note that the

Page 104: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.3 Results and Discussion: FTIR-Raman Spectroscopy

84

pressure-temperature conditions in the study by Schmidt72 were further away from

conditions of salt precipitation. The solutions investigated by Schmidt72 were also saturated

with quartz, and consequently the Raman spectra exhibited features related to associated

species of silicic acid plus sulfate. The present experiments were conducted in a silica free

environment, such that this effect can be excluded.

Figure 6-7: Raman spectra showing the ν1 stretching mode of the sulfate anion at 316°C and pressures of 110, 120, 130, 150, 180 and 220 bar.

Page 105: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

85

Figure 6-8: Pressure dependence of the relative peak areas of the ca. 973 and 993 cm-1

Raman peaks in a 10 wt.% Na2SO4 solution at 316°C. The variation in relative peak areas qualitatively reflects the relative abundances of free (unassociated) ions and ion pairs versus small and medium sized clusters.

6.3.2. Li2SO4-H2O

Lithium sulfate is (like sodium sulfate) a type 2 salt. We investigated solutions at a

concentration of 10 wt.% at a pressure of 95 bar (Figure 6-9). No additional signals were

observed up to a temperature of 297°C. A cubic fit to solubility data from literature137

predicts a solid precipitation temperature of ~298 °C. Our measurement at 298 °C revealed

additional signals at 1009 and 1120 cm-1, which are according to literature138 the ν1 and ν3

bands of the monoclinic solid phase. No clustering prior to the solid precipitation could be

identified from our spectra. This behavior is contrary to the one observed for Na2SO4 and

Na2CO3 solutions, but could be caused by the clustering happening only over a very small

temperature range. Another possibility is that the precipitation occurs via the formation of

Li2SO4∙H2O, which is the stable solid phase up to 232.8 °C.139

Page 106: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.3 Results and Discussion: FTIR-Raman Spectroscopy

86

Figure 6-9: Raman spectra of a 10 wt.% Li2SO4 solution at 95 bar and 297 and 298 °C.

6.3.1. Na2CO3-H2O

Measurements were performed on a 10 wt.% Na2CO3 solution. At ambient conditions the

symmetric stretching CO32- band was found at 1066 cm-1, which is in line with the findings

from Sipos et al.70 and from Frantz,74 who reports the band to appear at 1064-1065 cm-1. Our

measurements (Figure 6-10) at 288 °C and 85 bar showed that this band shifts to 1055 cm-1

and a signal of the HCO3- ion appears at 977 cm-1 which is in line with literature.71, 74 At

290 °C and 85 bar, a signal appears at 1074 cm-1, which can be attributed to clustered ions,

in analogy to the results of the Na2SO4 solutions. The carbonate mode in the solid Na2CO3

would appear at 1079 cm-1 and solid precipitation should set on at a temperature of 292 °C

according to interpolation from literature data,111, 120, 140 which supports our interpretation

of the 1074 cm-1 band being the one of clustered ions. In contrast to our results from the

Na2SO4 solutions, the clustering seems to appear even closer to the solubility limit and to

happen over an even smaller range of temperature, which might be caused by the

carbonate-bicarbonate equilibrium. Unfortunately the carbonate solutions exhibited a highly

corrosive behavior, leading to partial dissolution of the sapphire window and led to leakage

at the window sealing. This prevented us from further investigating these solutions.

9501'0001'0501'1001'150

Inte

nsi

ty /

a.u

.

Raman shift / cm-1

978 cm-1

free sulfate

1009 cm-1

solid Li2SO4

1120 cm-1

solid Li2SO4

297 °C

298 °C

Page 107: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

87

Figure 6-10: Raman spectra of a 10 wt.% Na2CO3 solution at 85 bar and 288 and 290 °C.

6.3.2. Na2HPO4-H2O

A 5 wt.% solution of Na2HPO4 was prepared and spectra were taken at 314°C, 120 bar. The

quality of the spectra was poor due to the rather low concentration of the salt, which led to

a low signal to noise ratio, which prevented an extensive analysis. The HPO42- band which is

located at 990 cm-1 at ambient conditions141 shifted to 980 cm-1 at the mentioned

conditions. According to literature122 a liquid immiscibility should occur at around these p-T

conditions. The spectra revealed no appearance of CIPs or other drastic changes under such

conditions. Marshall and Begun142 studied a 2 molal solution under isochoric conditions and

reported a shift of the hydrogen phosphate band from 989 cm-1 at ambient conditions to

984 cm-1 at 280 °C. Furthermore, they investigated the two liquid phases in equilibrium (as

Na2HPO4 is a type 1d salt) at 320 °C, where they found the phosphate band at 979 cm-1 in the

concentrated liquid and at 984 cm-1 in the dilute liquid. According to Marshall and Begun142

their results did not reveal the origin of the liquid immiscibility.

9509709901'0101'0301'0501'0701'0901'110

Ram

an in

ten

sity

/ a

.u.

Raman shift / cm-1

290 °C

997 cm-1

HCO3-

1055 cm-1

CO32-

1074 cm-1

clusters

288°C

Page 108: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.3 Results and Discussion: FTIR-Raman Spectroscopy

88

6.3.3. Na2SO4-K2SO4-H2O

Three ternary solutions were prepared at the same sulfate molality (0.7824 mol/kg H2O) as

in the experiments with 10 wt.% Na2SO4 solution, but replacing 1, 5 or 10 mol% of the

sodium with potassium. The measurements were conducted at 316 °C and 12 MPa (Figure

6-11). From the deconvolution of the signals at 973 and 993 cm-1 we see that the relative

peak area of the clusters decreases with increasing proportion of potassium in the solution

(i.e., increasing K+/Na+ ratio, Figure 6-12).

Figure 6-11: Raman spectra showing the ν1 stretching mode of the sulfate anion at 316°C and 120 bar at K+/(K+ + Na+) molar ratios of 0, 1, 5 and 10%.

The precipitation temperatures of the aforementioned solutions were determined using DSC

measurements. This analysis was done in order to test the hypothesis that the proportion of

associated ions (as revealed by spectroscopic measurements) is correlated to the

temperature difference between the experiment and the precipitation temperature. Stated

differently, if ion association leads to precipitation of solid, then the less associated solutions

(at equal temperatures) would be expected to correspond to those with higher precipitation

temperatures. The measurements were performed at isochoric conditions and therefore

H2O partitions into the gas phase during heating, leading to an increased salt concentration

in the remaining liquid phase. Hence, precipitation occurs at a temperature somewhat lower

Page 109: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

89

than expected based on the starting (bulk) composition. The precipitation temperature of

the 10 wt.% Na2SO4 solution at an average density of 300 kg/m3 is 316.6 °C, which is 3.7°C

lower than the literature data. We assumed that the temperature shift for the ternary

solutions will be approximately equal to this value for the binary solution, and therefore

applied a correction factor of +3.7 °C to the measured precipitation temperature for ternary

solutions (Table 6-1). The ΔT value is the shift in the precipitation temperature of the ternary

solutions relative to the pure sodium sulfate solution. Results indicate that as the proportion

of ionic clusters decreases (as observed from the spectroscopic measurements), the

temperature difference required to induce precipitation increases concomitantly. Thus, the

DSC measurements confirm that the solutions exhibiting more ion association at a given

temperature are those that are nearest in temperature to the three-phase curve. This result

suggests that ion association in sodium sulfate solutions correlates to the onset of

precipitation.

Figure 6-12: Dependence of the relative peak areas of the ca. 973 and 993 cm-1 Raman peaks in a 10 wt.% Na2SO4 solution at 316°C and 12 MPa on the K+/(K+ + Na+) molar ratio. The variation in relative peak areas qualitatively reflects the relative abundances of free (unassociated) ions versus contact ion pairs / clusters.

Page 110: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.4 Conclusions and Outlook

90

Table 6-1: Transition temperatures, T (V−L−S), as determined with HP-DSC and temperature differences, ΔT, for mixtures with an increasing content of potassium, n(K), compared to the pure Na2SO4 solution

n(K) /mol% T (V-L-S) /°C ΔT /°C

0 320.3 0.0

1 322.5 2.2

5 327.1 6.8

10 332.5 12.2

6.4. Conclusions and Outlook

The investigations using FT-IR Raman spectroscopy revealed that in Na2SO4 solutions clusters

are being formed prior to precipitation conditions. The amount of clusters increases with

conditions getting closer to precipitation conditions, meaning lower pressure and higher

temperature. Upon replacing some of the sodium with potassium less clusters are being

formed, which is also reflected in a higher precipitation temperature, as confirmed by

isochoric HP-DSC. The interpretation of the Raman spectra of the sodium sulfate solutions

was aided by Molecular Dynamics (MD) simulations (see next chapter), as the Raman spectra

do not reveal any direct structural information.

In solutions of Li2SO4 the formation of clusters was not observed. This might be related to

the formations of clusters happening over a very small temperature range, or the formation

of Li2SO4∙H2O as an intermediate during precipitation.

Solutions of Na2CO3 revealed a formation of clusters close to the p-T conditions at which the

solid should precipitate. Here the carbonate-bicarbonate equilibrium probably complicates

the system, which leads to the described behavior.

Unfortunately the used Raman spectrometer was only capable to perform measurements up

to 320° C. Above that temperature, the thermal background radiation became too strong,

which leads to saturating the detector.

Page 111: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 6

91

As a recommendation for further research the use of a Raman microscope in combination

with a capillary system can be suggested, similar to the one used by Wang et al.64 This would

allow also the investigation of different fluid phases in equilibrium with each other, such as

in the liquid immiscibility that occurs in the ternary Na-K-SO4-H2O system.

Another experimental problem was corrosion of the sapphire windows and gold gaskets of

the high-pressure cell, caused by the high electrolyte concentrations. This issue needs to be

addressed in the design of future spectroscopic setups, in order to ensure reliable operation

of such setups and also to guarantee a stable composition of the fluids under investigation.

Page 112: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

6.4 Conclusions and Outlook

92

Page 113: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

93

Studies on the Speciation in Hydrothermal Salt Solutions Chapter 7

with Molecular Dynamics Simulations

Parts of this chapter have been published as:

J. Reimer, M. Steele-MacInnis, J.M. Wambach, F. Vogel, Ion Association in Hydrothermal

Sulfate Solutions, studied by Modulated FT-IR-Raman and Molecular Dynamics, J. Phys.

Chem. B, 2015, 19 (30), 9847–9857. Used with permission. Copyright 2015 American

Chemical Society.

Furthermore, parts are considered for publication as:

J. Reimer, M. Steele-MacInnis, F. Vogel, Speciation and Structural Properties of Hydrothermal

Na-K-SO4 Solutions Studied by Molecular Dynamics Simulations, in preparation.

7.1. Introduction

Molecular dynamics simulations (MD) offer an important insight into the microscopic

behavior of molecular systems. Usually the sets of parameters for ions and water are

developed for simulations at ambient conditions, as the models were developed primarily

for simulating bio-molecules such as proteins and biological membranes. Nevertheless some

of the water models reproduce the pVT-properties of water at high temperatures quite

well.143 Also the force field parameters for ions are usually applicable at high temperatures,

keeping in mind that the results of such simulations will yield semi-quantitative results. In

this part of the thesis, MD-simulations were performed to support our experimental findings

from Raman-spectroscopy.

A comment on the nomenclature used in this chapter: In literature, the term "ion-pair" is

usually used to describe aggregates of two to four ions, whereas the term "cluster" is

commonly used to denote a conglomeration of three or more atoms or ions. Some overlap

thus occurs in the standard usage of these terms. In the present study, we will refer to

aggregates of up to three ions as ion-pairs, and species of more than three ions will be

Page 114: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.2 Methods: MD-Simulations

94

referred to as clusters. Furthermore, for convenience we will refer to clusters with up to

three sulfate ions as "small" clusters, those with six or more sulfate ions as "large" clusters,

and those of intermediate size (4 or 5 sulfate ions) as "medium" clusters. These designations

based on the number of anions in the cluster are clearly somewhat arbitrary, but aid in

concisely describing qualitative trends in the observations. The divisions between small,

medium and large clusters used here also have some basis in the geometry of the species:

small clusters (≤ 3 sulfate ions) commonly exhibit various spherical to chain-like geometries,

whereas large clusters (≥ 6 sulfate ions) may exhibit a structure approaching that of the

crystalline solid, as the 6 anions can fully separate a sodium cation from the surrounding

water molecules. These terms will be used to describe the speciation hereafter.

7.2. Methods: MD-Simulations

The MD simulations were conducted using the GROMOS MD-software144-147 with a modified

version of the 54a8-forcefield,148 incorporating the potassium ion parameters derived by Reif

and Hünenberger.149 Models for the sodium and sulfate ions are included in this forcefield

version as well as the SPC/E150 model for the water. This water model was chosen because of

its ability to reproduce the pVT and static permittivity properties of real water at high

temperatures.90 We used the NVT ensemble with 42 SO42– and 84 Na+ ions (or K+ ions, or

Na++K+ ions in different proportions; see Table 7-2) as well as 3000 water molecules in a

cubic box with periodic boundary conditions. The sulfate ions are tetrahedra with S-O bond

lengths of 0.15 nm, a partial charge of 0.54 e at the sulfur atom and partial charges

of -0.635 e at the oxygen atoms. The density was set to the experimental values from

Khaibullin et al.,104 for temperatures above 320 °C we used extrapolated values. The ternary

simulations were carried out at the same molar volume, by adjusting the density

accordingly. After an energy minimization of the simulation box at 0 K, the temperature was

increased to the desired value and the system was equilibrated for 100 ps. The solute and

solvent molecules were coupled to Berendsen-type thermostats151 with a relaxation time of

0.1 ps. Bond lengths and angles were constrained using the SHAKE algorithm152 with a

relative geometric tolerance of 10-4 for each timestep. The cutoff distances for the short and

long range electrostatic forces were set to 0.8 and 1.4 nm respectively. The static

permittivity of the reaction field was set to the values of pure liquid water at vapor

Page 115: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

95

saturation and the respective temperature.6 For the temperatures above the critical

temperature of pure water, the values along the critical isochore of pure water were used.

Production simulations were run for 6-8 ns of total simulation time (except the 25 °C

simulation, which lasted for 40 ns total simulation time), using a timestep of 2 fs. The

trajectories were written every 100 timesteps, i.e. 0.2ps. The simulations were run on the

Brutus cluster at ETHZ.

Production simulations were analyzed by calculating radial distribution functions as well as

time series of the interatomic distances between all sodium and sulfur atoms. These

interatomic-distance time series were then evaluated by calculating the respective ion-

pairs/clusters for each timestep of the simulation. The lifetimes of ionic species and their

relative abundances (expressed as the percentage of all ions in a cluster of given

composition, averaged over the total simulation time) were determined from the time-series

data.

Ion pairing and clustering were assessed based on the time-series interatomic (Na-S)

distances between all pairs of Na+ and SO42– ions. Ion association was considered to occur if

the Na-S interatomic distance for a given pair was less than or equal to a threshold distance

characterizing associated sodium sulfate. The threshold value characterizing associated

sodium sulfate at each temperature and pressure was estimated from the first minimum in

the Na-S radial distribution function. We performed a sensitivity analysis to assess the

degree to which the calculated speciation varies with the estimated threshold distance. We

found that the lifetimes of all clusters exhibit a maximum when the cutoff distance

corresponds to the first minimum of the corresponding radial distribution function. A change

of the cutoff distance by ± 0.01 nm around the global minimum in the corresponding radial

distribution function changes the relative abundance of free ions by approximately half a

percentage point, although the lifetime of the NaSO41- cluster is more strongly affected (up

to ±0.25 ps). Therefore the distance of the first minimum in the Na-S radial distribution

function was used to define associated ions hereafter.

Page 116: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.2 Methods: MD-Simulations

96

Table 7-1: Lennard-Jones parameters and partial charges used for the MD-simulations.

Atom / Ion C6 / kJ∙nm6 C12 / kJ∙nm12 Charge / e

S 9.984E-03 1.308E-5 +0.54

O (sulfate) 2.262E-03 1.415E-7 -0.635

Na+ 7.884E-5 7.290E-8 +1

K+ 1.296E-03 1.866E-06 +1

O (SPC/E-water) 2.617E-03 2.634E-06 -0.8476

H (SPC/E-water) 0 0 +0.4238

Table 7-2: Compositions, densities and reaction-field permittivities used in the MD-simulations

Composition Na84(SO4)42 Na63K21(SO4)42 Na42K42(SO4)42 Na21K63(SO4)42 K84(SO4)42 εRF

T / °C Density / kg/m3

25 1089.0 1095.1 1101.3 1107.4 1113.6 78.5 220 949.0 954.4 959.7 965.1 970.4 31.5 260 898.0 903.1 908.1 913.2 918.3 25.6 280 870.0 874.9 879.8 884.7 889.6 22.8 300 840.0 844.7 849.5 854.2 859.0 20.1 320 808.0 812.6 817.1 821.7 826.2 17.4 340 - - - - 791.7 14.6 360 - - - - 755.2 11.3 380 - - - - 716.8 5.3 400 - - 669.1 - 676.5 5.1 420 - - - - 634.3 5.0

Page 117: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

97

7.3. The Cluster Analysis Program

To analyze the trajectory files obtained from the MD-simulations, we wrote an object-

oriented computer program, using FORTRAN 90 as programming language, which evaluates

cluster formation by geometric means. In this chapter the principles of operation will be

briefly explained. GROMOS offers the possibility to produce time series data files, which

contain interatomic distances between one specific atom and a group of other atoms over

the whole trajectory. We used this tool to create one file for each sulfur atom in the

simulation, which includes the distances to all sodium ions over the simulated time. Those

files are read into our cluster analysis program and combined to form a three-dimensional

matrix, which then contains all Na-S distances for all timesteps (subroutine

data_conversion).

From this matrix the program creates a topology matrix for each timestep, based on the cut-

off distance provided by the user. The topology matrix (called mindist in the program code)

contains those pairs of sulfur and sodium atoms whose interatomic distance is less or equal

to the cut-off distance.

The topology matrix is then analyzed for each timestep by a consecutive algorithm which

searches the neighbors of all sulfur atoms and sodium atoms belonging to one cluster. After

completing one cluster, the program deletes all atoms belonging to this cluster from the

topology matrix and continues with searching for the next cluster until the topology file for

the given timestep is empty. This procedure avoids double-counting of clusters. The clusters

and their compositions (i.e. the identification numbers of sulfur and sodium atoms in the

cluster) are written to another matrix (which is four-dimensional) and a file.

The program continues with the analysis of the lifetimes and relative abundances of the

clusters. This is done by analyzing the matrix described above. The subroutine

lifetime_analysis follows the trace of each individual cluster in the four-dimensional cl_time

matrix, and creates a matrix consisting of four columns. The first and second column

contains the number of sulfate and sodium ions in the cluster, respectively. The third and

fourth column contains the consecutive number of the starting and ending timestep, i.e. the

timestep at which the cluster appears and the one when it disappears.

Page 118: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

98

Another matrix is then created, in which is counted how often a cluster of a certain

composition appears over the simulated time. From the difference between starting and

ending timestep and the step-size provided by the user, the average lifetime for the

different clusters is calculated.

In addition, the program calculates the relative abundance of the different clusters, given in

percentage of all ions which are in the cluster, averaged over the simulated time. This is

done by multiplying the cumulative lifetime of all clusters of a given composition with the

number of ions it comprises, and dividing by the total number of ions multiplied by the total

number of timesteps. This is done for all possible cluster compositions. The ‘residual’

percentage is assigned to the free ions, i.e. ions which are not part of any cluster/ion-pair.

Overall, the program is capable of performing the most important analysis of the trajectories

based on the time-series files. The program was written for the salt Na2SO4, but can easily be

modified for other salts/systems. Due to its object-oriented structure, further changes and

additions are also possible.

7.4. Results and Discussion: MD-Simulations

7.4.1. Simulations of the System Na2SO4-H2O

Simulation at room temperature (25 °C) reveals that essentially no clustering occurs in a

solution of 10 wt.% Na2SO4 (Figure 7-1). The proportion of free (unassociated) ions was as

high as 83.5% at these conditions. Wernersson and Jungwirth153 reported that in simulations

at room temperature, Na2SO4 tended to form solid-like clusters after 15-20 ns of simulated

time, when using the SPC/E water model. In our simulations no “precipitation” occurred at

room temperature over a simulated time of 40 ns. This difference might arise from two

reasons: Firstly, Wernersson and Jungwirth used a different sulfate models, which were

derived from Hartree-Fock calculations,154 with higher partial charges on the oxygen atoms

(-0.9 to -1.1 versus -0.635 in our simulations). Secondly, our simulations were conducted in

the NVT ensemble at experimentally determined densities, whereas Wernersson and

Jungwirth used the NpT ensemble at 1 atm pressure.

Page 119: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

99

The pressures in our simulations were calculated, but did not coincide with experimental

values, possibly owing to the low vapor pressure of the SPC/E water model compared to real

water,90 or the fact that the ion models are not optimized for high concentrations or

temperatures.

To assess qualitatively the variation in ion pairing and cluster formation in the simulations,

the clusters were arbitrarily categorized into three groups (Table 7-3) according to the small,

medium and large cluster nomenclature described in the introduction of this chapter: 1)

small clusters with up to three sulfate ions, whose abundance strongly increases with

increasing temperature from 25 to 260 °C and decreases with increasing temperature

thereafter (from 260 to 320 °C); 2) medium-sized clusters (Figure 7-2) containing four or five

sulfate ions, the abundance of which is maximized at 280 °C; and 3) large clusters with 6 or

more sulfate ions which occur in significant abundance only at temperatures greater than

280 °C, and become the dominant species at 300 and 320 °C. Comparison with experimental

solubility data suggests that those large clusters represent precipitated salt (Figure 7-7). This

result thus indicates that the onset of the precipitation occurs at ~280 °C in the simulations;

i.e. ~40 °C lower than in the experiment. This discrepancy between the saturation

temperature from MD versus experiment is perhaps not surprising, given that the ion

parameters were optimized for ambient conditions. Nevertheless, the trends in the MD

results consistently show that increasing temperature favors formation of associated ions

and higher clusters, as well as precipitation of solid salt.

Page 120: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

100

Figure 7-1: Simulation box at 25°C showing the predominance of fully dissociated ions. Sodium in blue, sulfate in beige, water molecules omitted. Visualized using VMD.155

Page 121: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

101

Tab

le 7

-3: P

aram

ete

rs a

nd

res

ult

s o

f th

e M

D s

imu

lati

on

s w

ith

th

e n

on

-po

lari

zab

le S

PC

/E w

ate

r m

od

el.

T / °C

D

en

sity

/

kg/m

3 ε 0

cu

t-o

ff

dis

tan

ce

/ n

m

fre

e io

ns

larg

est

clu

ste

r /t

ota

l nu

mb

er

of

ion

s p

er

clu

ste

r

mo

st a

bu

nd

ant

clu

ste

ra /

Na-

SO4-

abu

nd

ance

lon

gest

lif

etim

eb

/ N

a-SO

4-lif

etim

e/p

s

clu

ste

rsc

≤3

SO

42-

(s

mal

l)

4&

5 S

O4

2-

(med

ium

) ≥6

SO

42-

(l

arge

)

25

108

9

78

.5

0.4

25

8

3.5

%

7

1-1

-10

.3%

1

-1-3

.1

16

.5%

<<

0.1

%

0%

220

9

49

31

.5

0.4

45

29

.0%

4

5 1

-1-1

7.6

%

1-1

-2.8

5

9.5

%

6.6

%

4.9

%

260

8

98

25

.6

0.4

45

2

1.5

%

51

2-1

-13

.7%

1

-1-2

.7

60

.1%

1

0.6

%

7.9

%

280

8

70

22

.8

0.4

55

1

7.5

%

54

2-1

-13

.0%

1

-1-2

.6

55

.9%

1

1.6

%

15

.0%

300

8

40

20

.1

0.4

75

10

.5%

1

16

2-1

-6.2

%

1-1

-2.6

1

9.0

%

0.9

%

69

.6%

320

8

08

17

.4

0.4

75

8

.0%

1

19

2-1

-5.2

%

1-1

-3.0

1

2.6

%

0.1

%

79

.3%

a i.e

. th

e cl

ust

er w

ith

th

e h

igh

est

valu

e o

f th

e re

lati

ve a

bu

nd

ance

in

per

cen

t o

f al

l io

ns,

ave

rage

d o

ver

the

sim

ula

ted

tim

e. T

he

com

po

siti

on

of

the

clu

ster

is

give

n i

n n

um

ber

of

sod

ium

io

ns

– n

um

ber

of

sulf

ate

ion

s. T

hu

s 1

-1-1

0.3

% m

ean

ing

that

th

e cl

ust

er w

ith

the

hig

hes

t ab

un

dan

ce (

10

.3%

) h

as a

co

mp

osi

tio

n o

f N

a1(S

O4) 1

1- .

b i.

e. t

he

clu

ster

wit

h t

he

lon

gest

ave

rage

life

tim

e. C

om

po

siti

on

of

the

clu

ster

giv

en a

s in

a ; wit

h li

feti

me

in p

s.

c per

cen

tage

val

ues

ref

er t

o c

um

ula

ted

rel

ativ

e ab

un

dan

ces

in p

erce

nt

of

all i

on

s, a

vera

ged

ove

r th

e si

mu

late

d t

ime

Page 122: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

102

Figure 7-2: Simulation box at 260°C showing small- and medium-sized clusters. Sodium in blue, sulfate in beige, water molecules omitted. Visualized using VMD.155

The NaSO41- ion pair is the most abundant cluster below 260 °C. The average lifetime of this

species decreases from ~3.1 ps at room temperature to ~2.6 ps at 300°C; at 320°C the

lifetime is again ~3.0 ps. In contrast, in solutions of 1 molal NaCl, Driesner et al. 92 reported

average lifetimes of greater than 100 ps for ion pairs of up to 3 ions at room temperature,

and lifetimes of 13.1 ps for the neutral NaCl0 ion pair at 380 °C.

At and above 260 °C the Na2SO40 cluster becomes the dominant species and higher clusters

are also present at these conditions, which coincides with our observations from Raman

spectroscopy. In addition, the presence of ionic clusters is also in good agreement with the

results from Sharygin et al.88 for Li2SO4 solutions, in which neutral clusters up to 24 ions are

present at high temperatures.

Page 123: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

103

The abundance of free (unassociated) ions decreases almost linearly with increasing

temperature, from 82.7% at 25 °C to 9.5% at 320 °C. Over the same temperature range, the

proportion of ions occurring in larger clusters increases with increasing temperature. These

observations are consistent with results of Driesner et al.92 for NaCl solutions, which

indicated that higher clusters are favored at high temperatures compared to room

temperature, although NaCl exhibits significantly less tendency towards forming large

clusters.92

In the 300 °C simulation, almost all of the Na+ and SO42– ions assembled to form a single

large cluster, containing around 80 ions, during the first 2 ns of simulated time (Figure 7-3).

Over a longer simulation period (4 ns in total) this large cluster re-dissolved, yielding a

variety of ion pairs and clusters of smaller sizes. At 320 °C the salt clearly precipitated in the

simulation at around 2.4 ns of simulated time (Figure 7-4), indicated by the formation of a

single large cluster of Na+ and SO42– ions that did not re-dissolve.

Note that the relative abundances and lifetimes of larger clusters are relatively low

compared to smaller ion pairs, owing to the continuous exchange of ions between the large

cluster and the bulk fluid at the outside surface of the cluster. Stated differently, ions in the

peripheral parts of the cluster repeatedly associate and dissociate from the cluster over

short timescales, and consequently they are commonly recognized as individual ions or

smaller clusters in our speciation analysis.

The Na-S radial distribution functions (RDF) (Figure 7-5) show an increase in the first

maximum with increasing temperature, which represents a greater tendency towards ion

association. Interestingly, the increase is especially strong between 280 °C and 300 °C, where

the precipitation of the solid occurs. A similar increase of the first maximum of the RDF has

been reported by Reagan et al.93 in a study on cluster formation in dense NaCl solutions at

hydrothermal conditions.

Page 124: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

104

Figure 7-3: Simulation box at 300°C, showing the formation of a single large cluster which re-dissolves after 2.6 ns, as well as several medium-sized clusters. Sodium in blue, sulfate in beige, water molecules omitted. Visualized using VMD.155

Page 125: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

105

Figure 7-4: Simulation box at 320°C, showing the precipitated salt after 4 ns simulated time. Visualized using VMD.155

The position of the main maximum is approximately constant at 0.305 nm for all

temperatures. The RDF of the simulations at 300 °C and 320 °C furthermore exhibits a

strong, broad second maximum at 0.625 nm, which suggests that larger clusters with a

second coordination shell of ions exist. Alongside the global maximum (i.e., the first

maximum), a shoulder occurs at 0.385 nm for the simulations at 220-280 °C and becomes a

local maximum at 300 and 320 °C. This feature appears to originate from the different

coordination possibilities (mono- vs. bidentate) between Na+ and SO42– ions, or different

coordination positions in clusters with structures similar to the solid phase. Thus, the

equilibrium Na-S distance in associated ion pairs and clusters appears to be either 0.305 or

0.385 nm, depending on the local structure, and the average Na-S distance (~0.345 nm) is

close to the Na-S distance in solid sodium sulfate (0.342 nm) as well as aqueous NaSO4– ion

Page 126: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

106

pairs (0.345 nm).77 The global minimum of the RDF shifts from 0.425 nm at room

temperature to 0.475 nm at 320 °C, which may reflect the higher kinetic energy of the ions

at higher temperatures.

Average coordination numbers can be obtained by integrating the RDF from the origin to the

first minimum, and converting to number of ions based on the particle number density of

the fluid. The Na-S RDF yields an average coordination of 0.11 sulfate ions around a sodium

ion at room temperature. This value increases with increasing temperature, to 1.07 at

280 °C, which is approximately one sixth the coordination number in solid Na2SO4 (in which

each sodium ion is coordinated by 6 or 7 sulfate ions, depending on the solid phase that is

precipitated).156, 157 A higher proportion of small- to medium-sized clusters is responsible for

this increase. The coordination number further increases to 2.06 at 300 °C, and 2.62 at

320 °C, which reflects the precipitation at 300 °C and the precipitated solid itself at 320 °C.

The value at 320 °C is remarkably lower than the value expected for the solid. For individual

sodium ions the coordination number of surrounding sodium ions can rise up to ~4.8, which

reflects the fact that the system size is relatively small, hence a large proportion of the ions

of the precipitated solid salt are on the surface of the cluster, having less neighbors than the

ions in a bulk solid. In addition, some ions are still present in solution as smaller clusters,

causing the RDFs to differ from those expected for a solid salt. The differences in speciation

between dissolved versus precipitated salt are clearly indicated in a histogram of cluster

sizes at 220 °C and 320°C (Figure 7-6). This graph also suggests that small and medium sized

clusters (i.e., up to 5 sulfate ions) are soluble. The presence of (rare) clusters with 6 and

more sulfate ions at 220°C is due to the high abundance of small and medium sized clusters,

because of occasional collisions between these species during the MD simulation. This is also

reflected in the small lifetime of such larger clusters, which is typically between <0.4 ps (i.e.

just 1-2 timesteps in the MD trajectories), whereas small and medium sized clusters exhibit

average lifetimes >0.6 ps.

Page 127: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

107

Figure 7-5: Na-S radial distribution functions for the simulations with SPC/E water at different temperatures.

Figure 7-6: Cluster size based on the number of sulfate ions per cluster from MD simulations at 220°C and 300°C. Free sulfate ions are omitted.

0

10

20

30

40

50

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

g (

r )

r / nm

25°C

220°C

260°C

280°C

300°C

320°C

Page 128: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

108

Figure 7-7: Comparison of speciation (sum of free ions, small and medium sized clusters vs. larger clusters) with experimental solubility (dissolved salt vs. precipitated salt), showing that the precipitation of the solid salt occurs at ~40 °C lower than in the experiment. Experimental values from literature.102, 104-106, 110, 111, 117 Discrepancies between experiment and simulations may be attributed to the arbitrary subdivision into small, medium and large clusters.

0%

20%

40%

60%

80%

100%

0 50 100 150 200 250 300 350 400

rela

tive

ab

un

dan

ce /

wt.

%

T / °C

sum of free,small andmedium clusters

large clusters

dissolved salt,experimental values

precipitated salt,experimental values

Page 129: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

109

Figure 7-8: Running coordination number (CN) of sodium with sulfur. The value at the first minimum in the RDF (Figure 7-5) represents the conventional CN.

Figure 7-9: Radial distribution functions (RDF) between sodium and oxygen of the sulfate ions.

0

0.5

1

1.5

2

2.5

3

0.25 0.3 0.35 0.4 0.45 0.5

coo

rdin

atio

n n

um

ber

r / nm

coordination number of Na with S(sulfate)

320°C

300°C

280°C

260°C

220°C

25°C

0

10

20

30

40

50

60

70

80

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

g (

r )

r / nm

Na-O(sulfate) RDF

320°C 300°C

280°C 260°C

220°C 25°C

Page 130: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

110

Figure 7-10: Running coordination number of sodium with oxygen of the sulfate ions. The value at the first minimum in the RDF (Figure 7-9) represents the conventional CN.

Figure 7-11: Radial distribution functions (RDF) between sodium and oxygen of the water molecules.

0

1

2

3

4

5

6

0.2 0.25 0.3 0.35 0.4

coo

rdin

atio

n n

um

ber

r / nm

coordination number of Na with O(sulfate)

320°C300°C280°C260°C220°C25°C

0

1

2

3

4

5

6

7

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

g (

r )

r / nm

Na-O(water) RDF

25°C 220°C

260°C 280°C

300°C 320°C

Page 131: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

111

Figure 7-12: Running coordination number of sodium with oxygen of the water molecules. The value at the first minimum in the RDF (Figure 7-11) represents the conventional CN.

7.4.2. Simulations of the System NaxK2-xSO4-H2O

In Figure 7-13 to 7-15 the speciation based on the aforementioned subdivision into small,

medium and large clusters is shown for the binary solutions of Na2SO4, K2SO4, and for an

equimolar KNaSO4 solution. Precipitation of solid salt is signified by a drastic increase in the

proportion of ions bound in large clusters.129 Onset of salt precipitation occurs in the Na2SO4

solution at about 280°C (Figure 7-13), which is ~40°C below the experimental value, as

reported previously.129 The pure K2SO4 solution shows the onset of the precipitation at

around 360 °C (Figure 7-14), which is close to the experimental value (355 °C).117 The

underestimated precipitation temperature for the Na2SO4 solutions and the unexpectedly

gradual decrease in solubility with temperature for the K2SO4 solutions (e.g., the K2SO4

solubility should decrease dramatically towards the H2O critical point at ~374 °C) highlight

some drawback of classical MD simulations. Nevertheless, the overall trends in temperature-

solubility relations for sodium- versus potassium-sulfate solutions are in overall agreement

with the known properties of these solutions. In the equimolar mixture, the proportion of

large clusters increases slightly starting around 300-320 °C; the simulation at 400°C shows

0

1

2

3

4

5

6

7

8

0.2 0.25 0.3 0.35 0.4

coo

rdin

atio

n n

um

ber

r / nm

coordination number of Na with O(water)

25°C

220°C

260°C

280°C

300°C

320°C

Page 132: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

112

precipitation of the salt, the onset of the precipitation being around 320 °C (Figure 7-15). It is

noteworthy that in terms of the speciation versus temperature trends in Figure 7-13 to 7-15,

the equimolar ternary mixture shows some similar features to both binary end-members; for

example, the equimolar mixture shows decreasing small and medium clusters coupled with

onset of abrupt increase in large clusters; the temperature range of onset of precipitation in

the ternary mixture is intermediate between those of the binary end-members (Figure 7-15).

Figure 7-13: Speciation in a Na2SO4 solution, based on the number of sulfate ions per cluster.

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

0 100 200 300

Na2SO4 @ SPC/E

free ions

small clusters

medium clusters

large clusters

Page 133: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

113

Figure 7-14: Speciation in a K2SO4 solution, based on the number of sulfate ions per cluster.

Figure 7-15: Speciation in an equimolar mixture of Na2SO4 and K2SO4, based on the number of sulfate ions per cluster.

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

0 100 200 300 400 500

K2SO4 @ SPC/E

free ions

small clusters

medium clusters

large clusters

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

0 100 200 300 400

T / °C

KNaSO4 @ SPC/E

free ions

small clusters

medium clusters

large clusters

Page 134: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

114

Figure 7-16: Speciation in the ternary solutions of Na-K-SO4. A: abundance free ions; B: small clusters with up to 3 sulfate ions; C: medium sized clusters with 4 or 5 sulfate ions; D: large sized clusters with more than 5 sulfate ions.

A complete overview of the cluster analysis for both end-member Na2SO4 solution and K2SO4

solution, as well as intermediate (ternary) mixtures of these two from 25 to 320 °C is

presented in Figure 7-16. The first noteworthy observation is that the fraction of ions in large

clusters at the highest temperatures (300 and 320 °C) is significantly greater in Na2SO4

solutions when compared to K2SO4 solutions (Figure 7-16 D). The presence of large clusters

at 320 °C in MD simulations of Na2SO4 solutions was previously described as evidence for

solid precipitation at these conditions.129 As expected, based on temperature-solubility

relations,129 K2SO4 solutions show significantly less tendency towards large cluster formation

at equivalent conditions (Figure 7-16 D). Meanwhile, K2SO4 solutions show somewhat

Page 135: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

115

greater percentage of ions in medium-sized clusters compared to Na2SO4 solutions (Figure

7-16 C), but are dominated by ion pairs and small clusters even at 320 °C (Figure 7-16 B).

In the ternary mixtures, the results of the cluster analysis of the trajectories indicate a

dramatic decrease in the proportion of ions in large clusters at high temperatures, when

potassium is added to a sodium sulfate solution (Figure 7-16 D). Furthermore, the solutions

exhibit an increasing proportion of ions in small clusters at 300 and 320 °C as potassium is

added (replacing Na+ with K+; Figure 7-16 B). The fraction of ions in medium sized clusters

increases with increasing molar fraction of sodium ions for the simulations at temperatures

below 300 °C. At and above this temperature, the curves exhibit a maximum between

X (Na/(Na+K), mole fraction) =0.50-0.75 (Figure 7-16 C). This phenomenon indicates that

mixtures of potassium- and sodium-sulfate, within a discrete range of cation ratios, are

especially favorable for the formation of medium-sized clusters. Stated differently, this

observation indicates an interaction between sodium and potassium, affecting the ion

association, discussed in more detail below.

The data shown in Figure 7-16 indicate that varying the Na:K ratio in an aqueous Na-K-SO4

solution can result in dramatic effects on the speciation of the electrolytes, according to the

degree of ion association and clustering. Specifically, addition of Na+ to a solution rich in

potassium sulfate results in gradual increase in the degree of ion association (in the studied

temperature range), whereas addition of K+ to a sodium-sulfate-rich fluid causes a rather

abrupt and non-linear decrease in the degree of ion clustering. This implies a salting-in

effect, which was also experimentally reported,119, 129 as the large clusters represent the

formation of a solid phase. Moreover, the proportion of medium-sized clusters exhibits a

maximum in solutions containing approximately equal fractions of sodium plus potassium. It

is known13, 114, 158 that liquid-liquid immiscibility is present, but metastable, in the phase

diagrams of H2O-Na2SO4 and H2O-K2SO4. Moreover, this metastable phenomenon is

stabilized in ternary mixtures of H2O-NaxK2-xSO4.117 In light of this observation, the present

results might be considered to indicate a correlation between the abundance of medium-

sized clusters, and the stabilization of liquid immiscibility. For example, this result may

suggest that liquid immiscibility is favored by the increased presence of the medium sized

clusters. However, we reiterate that binary Na2SO4 solutions, which show a preponderance

of medium-sized clusters over a finite temperature range,129 do not exhibit stable liquid

Page 136: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

116

immiscibility. Thus, the present results suggest that the characteristics of medium-sized

clusters in ternary NaxK2-xSO4 solutions may be particularly favorable for liquid immiscibility.

The visualized molecular trajectories provide a hint towards such a phenomenon: In the

ternary solutions with similar concentrations of sodium and potassium, the predominant

species appear to be sub-linear chains of ions, with systematic differences in the locations of

sodium versus potassium (Figure 7-17), described below. Note again that the subdivision

into small, medium and large clusters is based on geometrical considerations and is

therefore arbitrary (used mainly for convenience and brevity in describing the results).

Nevertheless, the observation of elongated clusters of sodium-potassium-sulfate (commonly

with K+ located in terminal positions) at intermediate compositions and temperatures is

pervasive and clear in the present results.

Raman spectroscopic studies64 on MgSO4 solutions have revealed that in the concentrated

liquid phase (the so-called "heavy liquid") formed via liquid immiscibility, clustering of ions

occurs to a greater extent than just simple ion-pairs and triple ions.28 Similar phenomena

were also suggested for other systems with liquid immiscibilities;142 hence, the present

findings from MD simulations are consistent with spectroscopic observations. Unfortunately,

the methods employed in the present study do not allow robust assessment of whether

phase separation (liquid-liquid immiscibility) occurs in the simulations, because "phases" do

not readily segregate (as gravity is not included), and phase boundaries are poorly- or un-

resolved in visualized trajectories. Recent studies on spherical particles of different sizes159

may provide the methods needed to visually resolve phase separation and liquid-liquid

immiscibility in Na-K-SO4 solutions, for future investigations.

Page 137: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

117

Figure 7-17: Visualization of the trajectory of the simulation with at Na:K ratio of 3:1 at 300 °C, showing polymeric clusters. Sodium ions in blue, potassium ions in turquois, sulfate ion atoms in beige. Visualized using VMD.155

The differing effects of Na+ versus K+ on ion-association, clustering, and sulfate-salt solubility

in aqueous fluids can be related to the different structural roles of these two cations in

solution, as the visualization of the trajectories reveals (Figure 7-17). The sodium ions form

bridges between different sulfate ions, whereas potassium ions occupy terminal positions in

the clusters and are seldom coordinated to more than one sulfate anion. The roles of sodium

and potassium are thus analogous to those of "network forming" versus "network

modifying" cations, respectively, in the parlance of silicate melts and glasses.160 Observation

of the visualized trajectories reveals that aggregated ions with potassium in "bridging"

positions tend to disintegrate more frequently than those with sodium in the bridging

position, yielding smaller clusters in the case of potassium-rich fluids. One reason for this

could be the higher ionic radius of the potassium ion, which leads to a weaker electrostatic

interaction with the sulfate ions.

The weaker potassium-sulfate interaction, relative to the sodium-sulfate interaction, is also

reflected in the average lifetimes of the clusters and the directionality of reactions involving

small ion pairs. The average lifetime of the contact ion pair MSO4- (M = Na,K) in the binary

solutions is 1.3-1.4 ps for the K2SO4 solutions and 2.7-3.1 ps in case of the Na2SO4 solutions.

In the simulations of the ternary solutions at 300°C the MSO4- (M = Na,K) ion pair in a Na:K =

Page 138: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

118

3:1 mixture has an average lifetime of 2.3 ps, in a Na:K = 1:3 mixture the lifetime is reduced

to 1.7 ps, and the absence of large clusters suggests that the disappearance of MSO4- ion

pairs occurs almost exclusively by decomposition to free ions. For the M2SO40 cluster the

average lifetime is reduced from 1.8 to 1.3 ps over the same compositional range, again

showing the increased unidirectionality (towards decomposition of clusters) in the more

potassium-rich solutions. This shows that the clusters containing potassium ions become

more labile as the molar fraction of potassium increases, meaning that an increased rate for

the decomposition of the clusters into smaller units shifts the equilibrium (i.e. the ratio

between the rates of the formation of a cluster and its rate of decomposition) towards

smaller clusters. Stated differently, the presence of more and smaller clusters is favored by

an easier decomposition of the clusters upon potassium addition.

Radial distribution functions (RDF), obtained from the simulations support the findings from

the trajectory analysis. In Figure 7-18 and Figure 7-19 the Na-S and K-S RDFs of the

equimolar mixture of Na2SO4 and K2SO4 are given. Both RDFs show similar features, with a

pronounced first maximum located at ~0.30 nm in the Na-S RDF versus ~0.35 nm in the K-S

RDF, and a broader second maximum centered around 0.55 to 0.6 nm. Note, however that

the magnitude of g(r) is very different between the two systems: For the first maximum in

the Na-S RDF at 320 °C, g(r) is ~45, whereas g(r) for the first maximum in the K-S RDF at

320 °C is ~14. This difference in the magnitude of the RDFs reflects the greater correlation of

sodium with sulfate, compared to potassium with sulfate.

Integration of the RDFs over the spherical shell up to the first minimum, normalized

according to the number density of the solution, yields the average coordination numbers of

the ions. In Table 7-4 the coordination numbers of the different cations with the sulfate

anion are given. The coordination number of the potassium ion is lower than that of the

sodium ion (except in the simulations at room temperature).

Interestingly, although both the sodium-sulfate and potassium-sulfate coordination numbers

increase with increasing temperature (Table 7-4), the ratio of coordination numbers

between sodium-sulfate and potassium-sulfate remains almost constant between 220 to

320°C. This nearly constant ratio of ~1.4-1.6 (Na-S/K-S) indicates that sodium has a greater

affinity towards ion-pair and cluster formation than potassium. Another evidence for this

Page 139: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

119

higher cluster forming tendency is that the increase in the cation-O(sulfate) coordination

number with a corresponding decrease in the cation-O(water) coordination number is much

more pronounced for the sodium ion compared to the potassium ions (Table 7-4). Stated

differently, the sodium-ion favors association with sulfate ions over hydration by water at

given temperatures, whereas the hydration shell of the potassium ion undergoes only small

changes (showing a greater affinity for hydration, compared to Na+). This also supports the

hypothesis that cluster formation is the first step towards solid precipitation, as the

experimental solubility limit of pure K2SO4 solutions is at higher temperatures compared to

the one of pure Na2SO4 solutions (see Reimer and Vogel117 and references therein).

Furthermore, this explains the affinity of potassium for occupying the terminal positions in

the clusters (Figure 7-17), as in those positions more water molecules can surround the ion.

Moreover the cation-cation RDFs (Figure 7-20) at 300°C in the simulations at Na:K = 1:1

reveal that there is some correlation between the positions of the sodium ions, which

reflects the higher tendency of the sodium to form clusters with the sulfate ions, as this

correlation is most likely mediated through coordination of more than one sodium ion to a

sulfate ion. In contrast to that the potassium-potassium RDF shows less correlation, owing to

the finding that the potassium ions occupy the terminal positions of the clusters, hence

being further apart from each other and movements being less correlated. The Na-K RDF

shows an intermediate amount of interaction, which also indicates that both kinds of cations

(Na+ and K+) can coordinate to the sulfate ion simultaneously.

Figure 7-18: Temperature dependence of the Na-S radial distribution functions at a 1:1 molar ratio of Na:K.

0

10

20

30

40

50

0.2 0.4 0.6 0.8 1

g (

r )

r / nm

Na-S (Na:K=1:1)

25°C

220°C

260°C

280°C

300°C

320°C

Page 140: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.4 Results and Discussion: MD-Simulations

120

Figure 7-19: Temperature dependence of the K-S radial distribution functions at a 1:1 molar ratio of Na:K.

Table 7-4: Coordination numbers and relative coordination numbers at a molar ratio Na:K = 1:1.

Average coordination

number T / °C

25°C 220°C 260°C 280°C 300°C 320°C

Na-S 0.12 0.83 1.04 1.05 1.23 1.26

K-S 0.17 0.56 0.66 0.73 0.78 0.85

Na-S/K-S 0.71 1.49 1.58 1.43 1.57 1.48

Na-O(sulfate) 0.17 1.97 2.44 2.64 2.89 3.25

K-O(sulfate) 0.29 1.08 1.38 1.50 1.61 1.76

Na-O(sulfate)/ Na-S

1.42 2.37 2.34 2.51 2.35 2.58

K-O(sulfate)/ K-S

1.69 1.94 2.08 2.04 2.06 2.07

Na-O(water) 5.7 4.4 4.1 3.9 3.8 3.5

K-O(water) 7.0 6.4 6.4 6.3 6.1 6.0

0

2

4

6

8

10

12

14

16

0.2 0.4 0.6 0.8 1

g (

r )

r / nm

K-S (Na:K=1:1)

25°C220°C260°C280°C300°C320°C

Page 141: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

121

Figure 7-20: Radial distribution functions for the cations in the simulations with Na:K = 1:1 at 300°C.

Taken together, these results provide hints into the structural roles of sodium and potassium

in aqueous sulfate solutions, which in turn yield insights into the molecular-scale

phenomena controlling the macroscopic phase behavior in these systems. The results also

imply that the macroscopic salting-in observed in mixed-salt solutions and the apparent

stabilization of liquid immiscibility in the ternary system do not result from simple

dilution/uncommon-ion effects. In the absence of cation-cation interaction effects, we

would expect linear mixing relations reflected in the proportions of species in the end-

member solutions, owing to dilution by the other end-member. Instead, we see that

addition of small fractions of potassium serves to rapidly break up the large ion clusters in

sodium-sulfate-rich solutions, whereas addition of small fractions of sodium induces only

gradual ion association changes in potassium-sulfate-rich fluids. Meanwhile, at 300 and

320 °C, clusters with 4-5 sulfate ions increase in abundance as either Na+ or K+ is added to a

fluid initially enriched in the other cation. This is caused by the breakdown of larger clusters,

when starting from the end-member sodium-sulfate solution. When starting from the

potassium-sulfate end-member, this effect can be attributed to the higher clustering

tendency of the sodium ions and their tendency to occupy bridging positions between the

sulfate ions. A schematic overview of this behavior is presented in Figure 7-21.

0

1

2

3

4

5

6

7

0.2 0.4 0.6 0.8 1

g (

r )

r / nm

Na-Na 300°C

K-K 300°C

Na-K 300°C

Page 142: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.5 Conclusions and Outlook

122

Figure 7-21: Scheme for the changes in speciation at ~300 °C. Sulfate in pink; sodium in red; potassium in green.

7.5. Conclusions and Outlook

The simulations with the non-polarizable SPC/E water model yield results in good agreement

with our data obtained from Raman spectroscopy. This is somewhat surprising, as the ion

models used in our simulations were not designed for high temperature simulations. In any

case one cannot expect that classical MD simulations will reproduce experimental data,

especially so far from temperatures for which the models for water and ions are designed

and tested for. Furthermore, other parameters of the forcefield such as the static

permittivity need to be estimated so far. In our case the permittivity of pure liquid water

along the saturation curve at the respective temperature proved to be a good estimation,

even if real salt solution will most likely have higher values due to the influence of the ions.

We find that in the type 2 salt solution of sodium sulfate, ion association and clustering

occurs at conditions approaching the three-phase line. This observation is corroborated both

by Raman spectroscopic measurements (which show increasing contributions at ca.

993 cm-1, representing associated sulfate at high temperature), and MD simulations (which

show increasing proportions of associated ions and clusters at high temperature). In general,

the formation of ion pairs and clusters is favored at high temperatures because the

decreasing density of the fluid and the higher kinetic energy of the ions and water molecules

weakens the hydration of the ions, and hence the Coulomb interaction between the ions will

be less shielded by water molecules, leading to the aggregation of the ions.

From Raman spectroscopy, we find that the concentration of associated ions in the liquid

increases when precipitation conditions are approached, either by increasing temperature or

decreasing pressure. These results indicate that association of ions in solution precedes the

Page 143: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

123

precipitation of type 2 salts at high temperatures and may suggest an essential first step

towards precipitation. Molecular simulations further confirm that ions in sodium-sulfate

solutions tend to assemble into ion pairs and clusters as the saturation temperature is

approached, providing good qualitative explanation of the trends observed by spectroscopic

measurements. Clusters of ions (as observed in MD simulations) seem to represent the

precursors of either pre-nucleation clusters or critical nuclei that herald the transition from

single-phase liquid to the liquid-solid equilibrium. Taken together, the spectroscopic results

and molecular simulations provide a consistent picture of increasing ion association towards

the conditions where sodium sulfate precipitates. Based on these results, we suggest that

the macroscopic phase behavior of salt solutions may be correlated to the stability of the

aqueous species. As mentioned above, Na2SO4 solutions exhibit type 2 phase behavior, with

decreasing solubility at high temperatures. In contrast, NaCl solutions have a phase behavior

of type 1, in which the solubility increases with temperature up to the NaCl triple point. In

the present study, we find a decrease in the proportions of small clusters (of only a few ions)

and a concomitant increase in the proportion of medium- to large-sized clusters close to the

solubility limit of Na2SO4 solutions. The previous study of Driesner et al.92 investigated

speciation of sodium chloride solutions at similar pressure-temperature conditions and

reported that small, long-lived clusters of low charge dominated at high temperatures.

Comparing the present results for sodium sulfate (a type 2 salt) with the previous results92

for sodium chloride (a type 1 salt), we infer that the speciation of the sodium chloride in

high-temperature aqueous solutions is more favorable for retaining the salt in solution.

Specifically, the predominance of small clusters (ion pairs) with neutral or low charges in

sodium chloride solutions permits maintaining a high solubility of NaCl in high-temperature

water (which is a weak dielectric medium at these conditions) hence indicating phase

behavior of type 1. In contrast, the stability of small clusters in sodium sulfate solutions

decreases at high temperatures, and ions assemble into larger species that precede and lead

to the formation of large clusters representing the precipitated solid phase, hence signifying

phase behavior of type 2. A possible reason for this difference in the properties of the two

salt solutions is the difference in the charge of the anion. This suggestion is also supported

by the observation that NaHSO4 behaves as at type 1 salt161 because the HSO4- anion, being

isoelectronic but less charged compared to SO42-, is stabilized in hydrothermal solutions.68, 89

Page 144: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.5 Conclusions and Outlook

124

The Raman spectroscopic measurements showed that replacing sodium with potassium ions

decreases the proportion of clusters, and our DSC experiments reveal that this effect reflects

an increasing precipitation temperature with increasing potassium content. This salting-in

effect was studied by MD simulations in the ternary Na-K-SO4-H2O system.

Our results reveal that the replacement of sodium with potassium ions in hydrothermal salt

solutions leads to a drastic decrease in the tendency towards formation of large clusters. The

experimentally found increasing solubility can be attributed to this effect as the formation of

large clusters in the simulation represents the precipitation of solid salt. The abundance of

medium-sized clusters exhibits a maximum at equimolar mixtures at temperatures of 300

and 320 °C, and visual observation of the molecular trajectories indicates that the medium-

sized clusters in such solutions are dominated by linearly-arranged chains of ions, with

sodium ions in bridging positions between the sulfates, versus potassium ions in terminal

positions. These results suggest that polymeric clusters may be key to the experimentally

observed liquid immiscibility, which occurs at similar conditions in ternary Na-K-SO4

solutions.

We could attribute the changes in the speciation to structural effects, caused by the lower

tendency of potassium towards cluster formation, which is related to the higher ionic radius,

and hence a weaker interaction with the sulfate and stronger tendency towards solvent-

stabilized (hydrated) species at high temperatures. This explains on a microscopic level the

reason for the macroscopic phase behavior, and the resulting favorable behavior of mixtures

in the salt separation in the SCWG process.

One way to overcome the problem of the classical MD simulations only yielding semi-

quantitative results is conducting ab initio MD simulations, where not only the classical

equations of motion of the molecules are solved, based on parametrized interaction

potentials, but rather those interactions are calculated quantum-mechanically. This leads to

a huge demand of computation power, up to now allowing the simulation of smaller systems

so far, which is unfavorable in our case where the statistics of ion-association play an

important role. Currently new methods are being developed for studying larger systems,162

which possibly also allow more precise simulations at higher temperatures and pressures,

but up to now the biggest simulated systems have a size of around 1000 atoms and the

Page 145: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 7

125

simulated times are around 1 ns, which is not enough to provide good statistics for ion-

association processes.

Page 146: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

7.5 Conclusions and Outlook

126

Page 147: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

127

Part C Design of a Salt Separator for the Hydrothermal

Gasification Of Biomass

Page 148: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

128

Page 149: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 8

129

Salt Separator Design Chapter 8

Parts of this chapter have been published as:

E. De Boni, J. Reimer, G. Peng, H. Zöhrer, F. Vogel Paul Scherrer Institut, Patent

EP2918549 A1

Or are considered for publication as:

J. Reimer, G. Peng, E. De Boni, S. Viereck, J. Breinl, F. Vogel, A Novel Salt Separator for the

Hydrothermal Gasification of Biomass, in preparation.

8.1. Introduction

In PSI’s hydrothermal gasification process the separation of inorganic salts, which are

constituents of the feed biomass, is an important step, to avoid damaging the catalyst and to

prevent blocking of the plant. Investigations by Schubert et al.47, 163-166 showed that

especially type 2 salt solutions are responsible for those issues. Furthermore, coking of the

organic constituents and tar formation are problematic for this process.167, 168

Up to now a dip-tube configuration has been used as a salt separator (Figure 8-2, left part).

The biomass was preheated to around 300°C and then injected through the axial dip-tube

directly into the hot upper part of the salt separation vessel which is heated to 400-450°C.

Concentrated salt brine is removed at the bottom of the reactor. A radial exit port situated

at the top of the reactor is used for feeding the organics/water mixture to the catalytic

reactor, where the methanation takes place. For some model solutions this configuration led

to separation efficiencies up to 80-92%, which is however not enough to protect the catalyst

from being influenced or even damaged.164

Therefore a new approach for the salt separation vessel was considered and tested. The

thoughts that led to a patent application are presented in this part of the thesis.

Page 150: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

8.2 Theoretical Considerations/Patent Application

130

8.2. Theoretical Considerations/Patent Application

The main idea behind the salt separation vessel is to make use of the unique properties of

hot compressed water. The solubility of organics increases dramatically4 with increasing

temperature due to the fact that the static permittivity of water decreases under such

conditions.169 For electrolytes, this leads to a reduced solubility (see Figure 8-1).

Figure 8-1: Isobaric solubility limits of various salts. Adapted with permission from Tester et al.170. Copyright 1993 American Chemical Society.

In the salt separator design used in previous studies in our group,47-49, 51, 52 the preheated

feed stream was injected into the upper hot zone of the reactor. This caused several

problems e.g. the precipitation of solid salts in the upper zone and also the formation of

coke. To overcome the problems related to the precipitation of solid salts, Müller168

proposed the usage of a longer dip-tube, to inject the feed solution below the upper hot part

of the reactor.

Our new design ‘inverts’ the old one (Figure 8-2, mid and right part), featuring a riser tube,

through which the cold feed is injected into the middle of the salt separation vessel. The

lower part of the salt separator is kept at temperatures below 374°C. At those conditions,

Page 151: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 8

131

the salts should stay dissolved, whereas towards the upper, hotter end of the reactor, the

liquid will start separating into a dense and a less dense phase and the organics will most

likely also rise to the less dense phase and being dissolved there, as the water reaches

temperatures above its critical point. The lower density phase, consisting mainly of water

and organics, being depleted in salts, will further rise to the top zone of the vessel and can

be withdrawn at the top of the reactor.

Figure 8-2: Scheme of the KONTI-2 salt separator with dip- and riser-tube configuration and the KONTI-C salt separator (not drawn to scale).

This idea can be visualized conceptually with an isobaric cross section through the phase

diagram of a type 1a salt (Figure 8-3). A rise in temperature (red arrow in Figure 8-3), which a

salt solution would undergo upon injection into the salt separator, will result in the

formation of a liquid phase (blue arrow in Figure 8-3), being enriched in the salt, and a vapor

phase (green arrow in Figure 8-3), being depleted in salt. The vapor phase will rise up in the

reactor, due to buoyancy, whereas the liquid phase will start to accumulate at the bottom.

For a type 2 salt (Figure 8-4) the precipitation of the solid salt will occur as the temperature

crosses the solubility line (red arrow in Figure 8-4). As the solubility is highly related to the

static permittivity and hence also to the density of the fluid, this precipitation will most likely

Page 152: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

8.2 Theoretical Considerations/Patent Application

132

happen at the "pseudo interface" between sub- and supercritical (in terms of pure water)

conditions, where the density of the fluid drops rapidly. The particles formed might then

drop to the subcritical zone of the vessel and (hopefully) be re-dissolved, whereas the almost

salt-free vapor phase (green arrow in Figure 8-4) will rise in the reactor. The use of a non-

preheated feed slurry/solution enhances the effect of having a lower temperature at the

bottom of the salt separator.

The new design should also enable a controllable zone in the mid to upper part of the

reactor, where the density of the fluid drops, at around the critical temperatures in terms of

pure water. This zone will most likely enable a good salt separation.

Figure 8-3: Sketch of an isobaric cross-section of a phase diagram of type 1a. F: one phase fluid; V: vapor phase; L: liquid phase; S: solid phase.

It should be noted that mixtures of different types of salt will most likely behave either as a

pseudo-type 1 or pseudo-type 2 salt solution and not as solutions of individual salts of which

the mixture consists. For a full description of the hydrothermal phase behavior of aqueous

salt solutions see introduction of this thesis (Chapter 2).

Page 153: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 8

133

Figure 8-4: Sketch of an isobaric cross-section of a phase diagram of type 2 above the pressure of the lower critical endpoint p. F: one phase fluid; S: solid phase.

8.3. Results

Initial tests on the performance of the salt separator were carried out in the scope of the

PhD thesis of G. Peng.124 The new configuration was tested with model solutions using the

KONTI-2 process demonstration unit and showed an increased salt separation and sulfur

recovery.

During a gasification campaign conducted at the ZHAW Wädenswil, Switzerland, using the

KONTI-C process demonstration unit,124 temperature data of the salt separator had been

recorded. A microalgae-water slurry with a dry-matter content of 6.5 wt.% served as feed.

During the course of the experiment the pressure was maintained at around 280 bar and the

set temperatures of the salt separator set to 450°C and 240°C for the upper and lower

heater, respectively. The temperature data was interpolated assuming axial symmetry of the

whole salt separator. Figure 8-5 shows the resulting color-coded temperature field, as well

as the profile of the pseudo-critical line for pure water at ~300 bars. From this figure it is

clear that the lower part of the reactor is truly subcritical, while the fluid at the top of the

reactor will be at conditions above the critical point of pure water. This is the desired

behavior, as the temperature gradient will most likely lead to a continuing desalination of

Page 154: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

8.3 Results

134

the fluid upon its rise in the reactor, whereas in the lower part most likely a dense brine can

be formed, as described above.

Unfortunately at higher concentrations of the feed and upon pressure fluctuations the

temperature field showed strong fluctuations, which eventually led to a “breakthrough” of

low temperature fluid towards the catalytic reactor. Furthermore, the salt separator was

filled with black solid residue after a time on stream of ca. 70 hours.

Figure 8-5: Interpolated, measured temperature field of the KONTI-C salt separator operated with a 6.5 wt.% (dry matter) microalgae slurry. Black line: 400°C isotherm, i.e. the pseudo-critical temperature of pure water at 280 bar. Circles: Experimental data points. Numbers: Reactor height in mm. Courtesy of S. Viereck.171

Page 155: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 8

135

8.4. Summary and Outlook

We introduced a novel salt separator design using a riser tube to separate and recover

inorganics from biomass in a hydrothermal gasification process. The design was tested using

two different scales of the reactor and model solutions as well as real biomass. For the

model solutions the separation performance was much better that for the original dip-tube

configuration. Even in the presence of a complex feed stream, i.e. microalgae, the

temperature field showed favorable behavior.

To provide a successful operation of the plant with real biomass the parameters for pressure

and temperature need further optimization according to different contents of organics in the

feed stream. Furthermore, additional research on the phase behavior of salt solutions

containing organics is needed to fully understand the salt separation and recovery at

hydrothermal conditions.

Page 156: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

8.4 Summary and Outlook

136

Page 157: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

137

Part D Conclusions

Page 158: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

138

Page 159: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 9

139

Summary, Conclusions and Prospects Chapter 9

9.1. Summary and Conclusions

Our investigations using isochoric, high-pressure differential scanning calorimetry

proved that the phase behavior of salt solutions at hydrothermal conditions can be

qualitatively assessed, as the various types of phase transitions (formation of a

second liquid, solid precipitation, gas-liquid homogenization) are reflected in

different signal shapes.

A quantitative analysis of the DSC data allows precise determination of solubilities,

provided that experimental data on the density of the phases in equilibrium is

available.

Furthermore, a liquid immiscibility was found in the ternary system Na-K-SO4-H2O,

which arises through the stabilization of the liquid immiscibility, which is metastable

in the binary systems (Na2SO4-H2O, K2SO4-H2O). The presence of this liquid

immiscibility was validated by visual observation in an optical cell as a

complementary method. Later, the structural effects in such mixed sulfate solutions

were investigated using MD simulations. The results showed a lower clustering

tendency for the potassium ions, compared to the sodium ions, which is reflected in

the higher solubility of K2SO4 at high temperature, and has its origin in the weaker K-

SO4 interaction as compared to the Na-SO4 interaction. The weaker interaction was

mainly attributed to the larger ionic radius of the potassium. In addition this weaker

interaction leads to the formation of chain like polymeric ion-clusters in the ternary

solutions, which might be key to the reported liquid immiscibility.

The DSC study on ternary solutions provides insight into the role of cations and

anions for the formation of a certain phase behavior. In summary, the same trends,

which are also valid for binary solutions were observed, meaning that smaller ionic

radii of ions favor type 2 behavior, whereas larger ions most likely exhibit type 1d

behavior.

Page 160: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

9.1 Summary and Conclusions

140

An exception is the carbonate anion, which in binary solutions shows type 1d

behavior in case of potassium as the cation and type 2 behavior in case of sodium as

cation, even if the carbonate has the smallest ionic radius of the studied anions. The

Molecular Dynamics simulations on sulfate solutions showed that the tendency

towards cluster formation responsible for type 2 behavior. This clustering tendency is

mainly driven by the strength of the anion-cation interaction. This leads to the

hypothesis that in polyatomic anions also the number of coordination sites plays a

role, as the sulfate anion offers more possibilities for bidentate coordination of a

cation, compared to the carbonate ion, just due to its higher symmetry.

The FT-IR Raman studies revealed the formation of clusters prior to the precipitation

of the solid salt in Na2SO4 solutions over a temperature range of several degrees

Celsius and several tens of bar pressure. It was suggested that these clusters are the

precursors for the formation of crystallites (classical nucleation theory) or of the

amorphous pre-nucleation clusters (non-classical nucleation theory). The

interpretation of the Raman data was aided by MD simulations, which revealed a

high clustering tendency for Na2SO4 solutions at high temperatures. Based on these

results a revised ion association/clustering/precipitation scheme is suggested (Figure

9-1).

Investigation on other type 2 systems (Li2SO4-H2O, Na2CO3-H2O), using Raman

spectroscopy did not reveal clearly the formation of clusters prior to precipitation.

This might be attributed to a smaller temperature range over which the clustering

occurs (Li2SO4-H2O) or in case of the sodium carbonate solution, the presence of the

carbonate/bicarbonate equilibrium, which most likely influences the cluster

formation, as well as the high pH of such solutions which leads to partial dissolution

of the sapphire windows and hence to a ternary system with unknown and most

likely more complex behavior.

Page 161: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 9

141

Based on the considerations on the phase behavior, a novel type of salt separator for

the hydrothermal gasification of biomass was suggested. The new design features a

riser tube for injecting the feed solution into the separator. First experiments showed

high separation efficiencies with model solutions and real biomass, which is most

likely achieved through a controlled zone where the transition from a high-density to

a low-density fluid takes place.

Figure 9-1: Proposed extended model for ion association to precipitation of the solid salt for type 2 salts at hydrothermal conditions.

+

+

++ - - -

-

fully solvated ions ions separated by a

double solvation

layer

ions separated by a

single solvation

layer

Contact ion pair formation

Precipitation of

solid salt

Formation of

clusters in solution

Page 162: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

9.2 Recommendations for Future Research

142

9.2. Recommendations for Future Research

Future investigations on the phase behavior of hydrothermal solutions using DSC

should comprise the influence of organics, in form of model compounds. Some

preliminary data was presented in Chapter 3. The organic model compounds should

include volatile organic compounds such as alcohols and furan derivatives, and

substances with higher melting/boiling points such as glycerol, phenol and cresol

derivatives. The aim should be to cover a range of substances which might arise from

biomass hydrolysis. Furthermore, organic compounds with hetero-atoms such as

amino acids or phospholipids can be used. Possibly the stability of the compounds at

elevated temperatures needs to be assessed first.

The microscopic structure of hydrothermal salt solutions could be investigated in

more detail in the future by spectroscopic methods, such as Raman spectroscopy,

X-ray Raman spectroscopy (XRS), X-ray diffraction and neutron diffraction (possibly

with isotope substitution).

o For studying phases in equilibrium, a confocal Raman microscope with a

capillary device, similar to the one used by Wang et al.,64 seems to be

advantageous, as it allows measuring spectra in each phase separately. It

should be taken care of that the capillary material is inert under the applied

conditions; hence sapphire or diamond could be used.

o XRS measurements are capable to study elements with low energetic K-edges.

For instance the water structure at elevated pressures and temperatures was

studied with XRS,172, 173 hence it should be possible to study also association

of ions containing Na, K, S, and P.

o X-ray diffraction has been used to study ion association at room temperature

and slightly elevated temperatures.77 Application to high pressures and

temperatures would need a suitable spectroscopic cell.

Page 163: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

Chapter 9

143

o Neutron diffraction with isotope substitution can yield information on the

water structure and also water-ion interactions (see 62 and literature therein).

A properly designed cell is nevertheless the key to successfully performing

such measurements.

In any case, molecular dynamics and/or Monte Carlo simulations will be needed to

interpret the data obtained by spectroscopic measurements. E.g., empirical structure

refinement (EPSR) can be used to interpret scattering data.174 Furthermore, ab-initio

molecular dynamics could be used to overcome the semi-quantitative nature of the

results of classical MD simulations; provided that computational power will allow the

simulation of lager systems and longer timescales in the future.

To further enhance the salt separation performance in the hydrothermal gasification

process, a salt separator setup equipped with various analytical devices (e.g. on-line

conductivity and density measurements) would be advantageous to yield deeper

insight into the characteristics of the process and would allow a better control and

optimization of the process parameters.

In general, basic research on the speciation and phase behavior of hydrothermal salt

solutions (especially of more complex systems and mixtures) will not only yield crucial

data for the salt separation step in SCWG processes, but will also be of interest for

geochemistry, geothermal power plants, and other processes where hot compressed

water is used as process medium.

Page 164: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

9.2 Recommendations for Future Research

144

Page 165: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

145

References

1. IEA, World Energy Outlook 2012. OECD Publishing: Paris.

2. Vogel, F., In Handbook of Green Chemistry Volume 2 Heterogeneous Catalysis, Anatas, P.; Crabtree, R., Eds. Wiley-VCH: Weinheim, 2009; Vol. 2, pp 281-324.

3. Waldner, M. H. Catalytic Hydrothermal Gasification of Biomass for the Production of Synthetic Natural Gas. PhD Thesis, ETH Zürich, 2007.

4. Rebert, C. J.; Kay, W. B., The Phase Behavior and Solubility Relations of the Benzene-Water System. AIChE Journal 1959, 5, 285-289.

5. Lemmon, E. W.; McLinden, M. O.; Friend, D. G., Thermophysical Properties of Fluid Systems. National Institute of Standards and Technology: Gaithersburg MD.

6. Fernández, D. P.; Goodwin, A. R. H.; Lemmon, E. W.; Sengers, J. M. H.; Williams, R. C., A Formulation for the Static Permittivity of Water and Steam at Temperatures from 238 K to 873 K at Pressures up to 1200 MPa, including Derivatives and Debye-Hückel Coefficients. Journal of Physical and Chemical Reference Data 1997, 26, 1125-1166.

7. Marshall, W. L.; Franck, E. U., Ion Product of Water Substance, 0-1000°C, 1-10,000 Bars New International Formulation and Its Background. J. Phys. Chem. Ref. Data 1981, 10, 295-304.

8. Yardley, B. W.; Bodnar, R. J., Fluids in the Continental Crust. Geochemical Perspectives 2014, 3, 127.

9. Garnsey, R., Corrosion of PWR Steam Generators. Nucl. Energy 1979, 18, 117-132.

10. Hydrothermal Experimental Data. John Wiley & Sons, Ltd.: Chichester, UK, 2008.

11. Scott, R. L.; van Konynenburg, P. H., Static Properties of Solutions. Van der Waals and Related Models for Hydrocarbon Mixtures. Discussions of the Faraday Society 1970, 49, 87-97.

12. Valyashko, V. M., Phase Equilibria in Binary and Ternary Hydrothermal Systems. In Hydrothermal Experimental Data, John Wiley & Sons, Ltd: 2008; pp 1-133.

13. Valyashko, V. M., Chapter 15 - Phase Equilibria of Water-Salt Systems at High Temperatures and Pressures. In Aqueous Systems at Elevated Temperatures and Pressures, Palmer, D. A.; Fernández-Prini, R.; Harvey, A. H., Eds. Academic Press: London, 2004; pp 597-641.

14. Driesner, T.; Heinrich, C. A., The System H2O–NaCl. Part I: Correlation Formulae for Phase Relations in Temperature–Pressure–Composition Space from 0 to 1000 °C, 0 to 5000 bar, and 0 to 1 XNaCl. Geochimica et Cosmochimica Acta 2007, 71, 4880-4901.

Page 166: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

146

15. Kennedy, G. C.; Wasserburg, G. J.; Heard, H. C.; Newton, R. C., The Upper Three-Phase Region in the System SiO2 -H2O. American Journal of Science 1962, 260, 501-521.

16. Valyashko, V. M., Fluid Phase Diagrams of Ternary Systems with One Volatile Component and Immiscibility in Two of the Constituent Binary Mixtures. PCCP Phys. Chem. Chem. Phys. 2002, 4, 1178-1189.

17. Valyashko, V. M., Solubility of Salts in the Presence of Vapor at 350°C, 370°C and 380°C in the K2SO4-KLiSO4-H2O System. Zhurnal Neorg. Khimii 1975, 20, 1129-1131.

18. Urusova, M. A.; Valyashko, V. M., Phase Equilibria and Critical Phenomena in the Li2SO4-K2SO4-H2O System at 380-400°C: Specifics of Ternary Phase Diagrams with Type 2 (p-Q) Boundary Subsystems. Russ. J. Inorg. Chem. 2012, 57, 579-591.

19. Urusova, M. A.; Valyashko, V. M., Supercritical Phase Equilibria in the Ternary Water-Salt System Na2CO3-K2CO3-H2O. Russ. J. Inorg. Chem. 2005, 50, 1754-1767.

20. Urusova, M. A.; Valyashko, V. M., Solubility Isotherm and Liquid-Liquid Phase Separation in the Na3PO4-Na2HPO4-H2O System at 350°C and under Vapor Pressure. Russ. J. Inorg. Chem. 2001, 46, 777-783.

21. Urusova, M. A.; Valyashko, V. M., Sodium Carbonate Solubility in the Na2CO3-NaCl-H2O System. Russ. J. Inorg. Chem. 2002, 47, 1581-1585.

22. Urusova, M. A.; Valyashko, V. M., Phase Equilibria and Heterogenization of Supercritical Fluids in the K2SO4-K2CO3-H2O System. Russ. J. Inorg. Chem. 2009, 54, 759-771.

23. Urusova, M. A.; Valyashko, V. M.; Grigor’ev, I. M., K2SO4-KCl-H2O Phase Diagram in the Area of Heterogenization of Homogeneous Supercritical Fluids. Russ. J. Inorg. Chem. 2007, 52, 405-418.

24. Urusova, M. A.; Valyashko, V. M., K2SO4-K2HPO4-H2O Phase Diagram in the Region of Heterogeneous Supercritical Fluids. Russ. J. Inorg. Chem. 2008, 53, 604-616.

25. Wagner, C., Theorie der Alterung von Niederschlägen durch Umlösen (Ostwald-Reifung). Zeitschrift für Elektrochemie 1961, 65, 581-591.

26. Ostwald, W., Über die vermeintliche Isomerie des roten und gelben Quecksilberoxyds und die Oberflächenspannung fester Körper. Zeitschrift für Physikalische Chemie 1900, 34, 495-503.

27. Lifshitz, I. M.; Slyozov, V. V., The Kinetics of Precipitation From Supersaturated Solid Solutions. Journal of Physics and Chemistry of Solids 1961, 19, 35-50.

28. Gebauer, D.; Cölfen, H., Prenucleation Clusters and Non-Classical Nucleation. Nano Today 2011, 6, 564-584.

29. Hodes, M.; Marrone, P. A.; Hong, G. T.; Smith, K. A.; Tester, J. W., Salt Precipitation and Scale Control in Supercritical Water Oxidation - Part A: Fundamentals and Research. Journal of Supercritical Fluids 2004, 29, 265-288.

Page 167: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

147

30. Marrone, P. A.; Hodes, M.; Smith, K. A.; Tester, J. W., Salt Precipitation and Scale Control in Supercritical Water Oxidation - Part B: Commercial/Full-Scale Applications. Journal of Supercritical Fluids 2004, 29, 289-312.

31. Hurst, W. S.; Hodes, M. S.; Bowers Jr, W. J.; Bean, V. E.; Maslar, J. E.; Griffith, P.; Smith, K. A., Optical Flow Cell and Apparatus for Solubility, Salt Deposition and Raman Spectroscopic Studies in Aqueous Solutions near the Water Critical Point. The Journal of Supercritical Fluids 2002, 22, 157-166.

32. Leusbrock, I.; Metz, S. J.; Rexwinkel, G.; Versteeg, G. F., The Solubilities of Phosphate and Sulfate Salts in Supercritical Water. Journal of Supercritical Fluids 2010, 54, 1-8.

33. Leusbrock, I.; Metz, S. J.; Rexwinkel, G.; Versteeg, G. F., The Solubility of Magnesium Chloride and Calcium Chloride in Near-critical and Supercritical Water. Journal of Supercritical Fluids 2010, 53, 17-24.

34. Akgul, G.; Kruse, A., Influence of Salts on the Subcritical Water-Gas Shift Reaction. Journal of Supercritical Fluids 2012, 66, 207-214.

35. Kruse, A.; Forchheim, D.; Gloede, M.; Ottinger, F.; Zimmermann, J., Brines in Supercritical Biomass Gasification: 1. Salt Extraction by Salts and the Influence on Glucose Conversion. Journal of Supercritical Fluids 2010, 53, 64-71.

36. Armellini, F. J.; Tester, J. W.; Hong, G. T., Precipitation of Sodium-Chloride and Sodium-Sulfate in Water from Sub- to Supercritical Conditions - 150 to 550°C, 100 to 300 bar. Journal of Supercritical Fluids 1994, 7, 147-158.

37. Armellini, F. J. Phase Equilibria and Precipitation Phenomena of Sodium Chloride and Sodium Sulfate in Sub- and Supercritical Water. PhD Thesis, Massachusetts Institute of Technology, 1993.

38. Armellini, F. J.; Tester, J. W., Solubility of Sodium-Chloride and Sulfate in Subcritical and Supercritical Water-Vapor from 450-500°C and 100-250 bar. Fluid Phase Equilibria 1993, 84, 123-142.

39. Hodes, M.; Griffith, P.; Smith, K. A.; Hurst, W. S.; Bowers, W. J.; Sako, K., Salt solubility and deposition in high temperature and pressure aqueous solutions. AIChE Journal 2004, 50, 2038-2049.

40. Rogak, S. N.; Teshima, P., Deposition of sodium sulfate in a heated flow of supercritical water. AIChE Journal 1999, 45, 240-247.

41. Lümmen, N.; Kvamme, B., Kinetics of NaCl Nucleation in Supercritical Water Investigated by Molecular Dynamics Simulations. PCCP Phys. Chem. Chem. Phys. 2007, 9, 3251-3260.

42. Nahtigal, I. G.; Zasetsky, A. Y.; Svishchev, I. M., Nucleation of NaCl Nanoparticles in Supercritical Water: Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2008, 112, 7537-7543.

Page 168: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

148

43. Svishchev, I. M.; Zasetsky, A. Y.; Nahtigal, I. G., Molecular Dynamic Simulations of Strontium Chloride Nanoparticle Nucleation in Supercritical Water. The Journal of Physical Chemistry C 2008, 112, 20181-20189.

44. Makaev, S. V.; Bitokhov, T. M.; Kravchuk, K. G.; Urusova, M. A.; Valyashko, V. M., Salt Deposition from Hydrothermal Solutions in a Flow Reactor. Russ. J. Phys. Chem. B 2011, 5, 1045-1055.

45. Marrone, P. A., Supercritical Water Oxidation-Current Status of Full-Scale Commercial Activity for Waste Destruction. Journal of Supercritical Fluids 2013, 79, 283-288.

46. Martin, A.; Bermejo, M. D.; Cocero, M. J., Recent Developments of Supercritical Water Oxidation: A Patents Review. Recent Patents on Chemical Engineering 2011, 4, 219-230.

47. Schubert, M.; Aubert, J.; Mueller, J. B.; Vogel, F., Continuous Salt Precipitation and Separation from Supercritical Water. Part 3: Interesting Effects in Processing Type 2 Salt Mixtures. Journal of Supercritical Fluids 2012, 61, 44-54.

48. Schubert, M.; Regler, J. W.; Vogel, F., Continuous Salt Precipitation and Separation from Supercritical Water. Part 1: Type 1 Salts. The Journal of Supercritical Fluids 2010, 52, 99-112.

49. Schubert, M.; Regler, J. W.; Vogel, F., Continuous Salt Precipitation and Separation from Supercritical Water. Part 2. Type 2 Salts and Mixtures of Two Salts. The Journal of Supercritical Fluids 2010, 52, 113-124.

50. Peterson, A. A.; Tester, J. W.; Vogel, F., Water-in-Water Tracer Studies of Supercritical-Water Reversing Jets Using Neutron Radiography. Journal of Supercritical Fluids 2010, 54, 250-257.

51. Peterson, A. A.; Vontobel, P.; Vogel, F.; Tester, J. W., Normal-Phase Dynamic Imaging of Supercritical-water Salt Precipitation Using Neutron Radiography. Journal of Supercritical Fluids 2009, 49, 71-78.

52. Peterson, A. A.; Vontobel, P.; Vogel, F.; Tester, J. W., In Situ Visualization of the Performance of a Supercritical-water Salt Separator Using Neutron Radiography. Journal of Supercritical Fluids 2008, 43, 490-499.

53. Bierrum, N., Die Dissoziation der Starken Elektrolyte. Zeitschrift für Elektrochemie und angewandte physikalische Chemie 1918, 24, 321-328.

54. Sutherland, W., Ionization, Ionic Velocities, and Atomic Sizes. Philosophical Magazine 1902, 3, 161-177.

55. Bjerrum, N., K. Dan. Vidensk. Selsk. 1926, 7.

56. Eigen, M.; Tamm, K., Schallabsorption in Elektrolytlösungen als Folge Chemischer Relaxation. 2. Messergebnisse und Relaxationsmechanismen für 2-2-Wertige Elektrolyte. Zeitschrift Für Elektrochemie 1962, 66, 107-121.

Page 169: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

149

57. Eigen, M.; Tamm, K., Schallabsorption in Elektrolytlösungen als Folge chemischer Relaxation. 1. Relaxationstheorie der Mehrstufigen Dissoziation. Zeitschrift Für Elektrochemie 1962, 66, 93-107.

58. Filimonova, Z. A.; Lileev, A. S.; Lyashchenko, A. K., The Complex Dielectric Constant and Relaxation in Aqueous Solutions of Alkali-Metal Nitrates. Russ. J. Inorg. Chem 2002, 47, 1890-1896.

59. Lyashchenko, A. K.; Kobelev, A. V.; Karataeva, I. M.; Lileev, A. S., Temperature Changes of Dielectric Permittivity and Relaxation in Aqueous Lithium Iodide Solutions. Russ. J. Inorg. Chem. 2014, 59, 757-765.

60. Maribo-Mogensen, B.; Kontogeorgis, G. M.; Thomsen, K., Modeling of Dielectric Properties of Aqueous Salt Solutions with an Equation of State. The Journal of Physical Chemistry B 2013, 117, 10523-10533.

61. Marcus, Y.; Hefter, G., Ion Pairing. Chem. Rev. 2006, 106, 4585-4621.

62. Driesner, T., The Molecular-Scale Fundament of Geothermal Fluid Thermodynamics. Reviews in Mineralogy and Geochemistry 2013, 76, 5-33.

63. Kohl, W. Ramanspektren Wässriger Medien bis 500°C und 3000 bar. Dissertation, Universität Karlsruhe, Karlsruhe, 1989.

64. Wang, X.; Chou, I. M.; Hu, W.; Burruss, R. C., In Situ Observations of Liquid–Liquid Phase Separation in Aqueous MgSO4 Solutions: Geological and Geochemical Implications. Geochimica et Cosmochimica Acta 2013, 103, 1-10.

65. Madekufamba, M.; Tremaine, P., Ion Association in Aqueous MgSO4 and NiSO4 from 25 to 275°C at 19 MPa by AC Conductance and Raman Spectroscopy. Geochimica Et Cosmochimica Acta 2010, 74, A654-A654.

66. Jahn, S.; Schmidt, C., Speciation in Aqueous MgSO4 Fluids at High Pressures and High Temperatures from ab Initio Molecular Dynamics and Raman Spectroscopy. The Journal of Physical Chemistry B 2010, 114, 15565-15572.

67. Rull, F.; Balarew, C.; Alvarez, J. L.; Sobron, F.; Rodriguez, A., Raman Spectroscopic Study of Ion Association in Aqueous Magnesium Sulphate Solutions. Journal of Raman Spectroscopy 1994, 25, 933-941.

68. Rudolph, W. W.; Brooker, M. H.; Tremaine, P., Raman- and Infrared Spectroscopic Investigation of Aqueous ZnSO4 Solutions from 8°C to 165°C: Inner- and Outer-Sphere Complexes. Z. Phys. Chemie-Int. J. Res. Phys. Chem. Chem. Phys. 1999, 209, 181-207.

69. Zhang, Y.-H.; Chan, C. K., Understanding the Hygroscopic Properties of Supersaturated Droplets of Metal and Ammonium Sulfate Solutions Using Raman Spectroscopy. J. Phys. Chem. A 2002, 106, 285-292.

Page 170: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

150

70. Sipos, P.; Bolden, L.; Hefter, G.; May, P. M., Raman Spectroscopic Study of Ion Pairing of Alkali Metal Ions with Carbonate and Sulfate in Aqueous Solutions. Australian Journal of Chemistry 2000, 53, 887-890.

71. Rudolph, W. W.; Irmer, G.; Konigsberger, E., Speciation Studies in Aqueous HCO3--

CO32- Solutions. A Combined Raman Spectroscopic and Thermodynamic Study. Dalton

Transactions 2008, 900-908.

72. Schmidt, C., Raman Spectroscopic Study of a H2O+Na2SO4 Solution at 21- 600 °C and 0.1 MPa to 1.1 GPa: Relative Differential ν1-SO4

2- Raman Scattering Cross Sections and Evidence of the Liquid-Liquid Transition. Geochimica et Cosmochimica Acta 2009, 73, 425-437.

73. Martinez, I.; Sanchez-Valle, C.; Daniel, I.; Reynard, B., High-Pressure and High-Temperature Raman Spectroscopy of Carbonate Ions in Aqueous Solution. Chemical Geology 2004, 207, 47-58.

74. Frantz, J. D., Raman Spectra of Potassium Carbonate and Bicarbonate Aqueous Fluids at Elevated Temperatures and Pressures: Comparison with Theoretical Simulations. Chemical Geology 1998, 152, 211-225.

75. Mantegazzi, D.; Sanchez-Valle, C.; Reusser, E.; Driesner, T., Thermodynamic Properties of Aqueous Sodium Sulfate Solutions to 773 K and 3 GPa Derived from Acoustic Velocity Measurements in the Diamond Anvil Cell. The Journal of Chemical Physics 2012, 137, 224501.

76. Matsumoto, Y.; Harada, H.; Yui, K.; Uchida, H.; Itatani, K.; Koda, S., Raman Spectroscopic Study of Aqueous Alkali Sulfate Solutions at High Temperature and Pressure to Yield Precipitation. The Journal of Supercritical Fluids 2009, 49, 303-309.

77. Xu, J.; Fang, Y.; Fang, C., Structure of Aqueous Sodium Sulfate Solutions Derived from X-Ray Diffraction. Chin. Sci. Bull. 2009, 54, 2022-2027.

78. Fulton, J. L.; Chen, Y.; Heald, S. M.; Balasubramanian, M., Hydration and Contact Ion Pairing of Ca2+ with Cl- in Supercritical Aqueous Solution. Journal of Chemical Physics 2006, 125, 094507.

79. Fulton, J. L.; Hoffmann, M. M.; Darab, J. G.; Palmer, B. J.; Stern, E. A., Copper(I) and Copper(II) Coordination Structure under Hydrothermal Conditions at 325°C: An X-Ray Absorption Fine Structure and Molecular Dynamics Study. Journal of Physical Chemistry A 2000, 104, 11651-11663.

80. Hoffmann, M. M.; Darab, J. G.; Palmer, B. J.; Fulton, J. L., A Transition in the Ni2+ Complex Structure from Six- to Four-Coordinate upon Formation of Ion Pair Species in Supercritical Water: An X-Ray Absorption Fine Structure, Near-Infrared, and Molecular Dynamics Study. Journal of Physical Chemistry A 1999, 103, 8471-8482.

81. Fulton, J. L.; Pfund, D. M.; Wallen, S. L.; Newville, M.; Stern, E. A.; Ma, Y. J., Rubidium Ion Hydration in Ambient and Supercritical Water. Journal of Chemical Physics 1996, 105, 2161-2166.

Page 171: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

151

82. Charnock, J. M.; Henderson, C. M. B.; Seward, T. M., High-Temperature Indium(III) Solutions. Journal of Synchrotron Radiation 1999, 6, 607-608.

83. Seward, T. M.; Henderson, C. M. B.; Suleimenov, M.; Charnock, J. M., X-ray Absorption Spectroscopic Studies of Halide Ion Hydration in Hydrothermal Solutions. Geochimica Et Cosmochimica Acta 2006, 70, A573-A573.

84. Seward, T. M.; Henderson, C. M. B.; Charnock, J. M.; Driesner, T., An EXAFS Study of Solvation and Ion Pairing in Aqueous Strontium Solutions to 300 Degrees C. Geochimica Et Cosmochimica Acta 1999, 63, 2409-2418.

85. Seward, T. M.; Henderson, C. M. B.; Charnock, J. M., Synchrotron X-ray Absorption (EXAFS) Measurements of Ion-Solvent Interaction in Aqueous Solutions up to 350 Degrees C. Abstracts of Papers of the American Chemical Society 1997, 214, 29-GEOC.

86. Seward, T. M.; Henderson, C. M. B.; Charnock, J. M.; Dobson, B. R., An X-ray Absorption (EXAFS) Spectroscopic Study of Aquated Ag+ in Hydrothermal Solutions to 350 Degrees C. Geochimica Et Cosmochimica Acta 1996, 60, 2273-2282.

87. Goralski, P.; Chabanel, M., Vibrational Study of Ionic Association in Aprotic Solvents. 11. Formation and Structure of Mixed Aggregates between Lithium Halides and Lithium Thiocyanate. Inorganic Chemistry 1987, 26, 2169-2171.

88. Sharygin, A. V.; Grafton, B. K.; Xiao, C.; Wood, R. H.; Balashov, V. N., Dissociation Constants and Speciation in Aqueous Li2SO4 and K2SO4 from Measurements of Electrical Conductance to 673K and 29MPa. Geochimica et Cosmochimica Acta 2006, 70, 5169-5182.

89. Hnedkovsky, L.; Wood, R. H.; Balashov, V. N., Electrical Conductances of Aqueous Na2SO4, H2SO4, and Their Mixtures:  Limiting Equivalent Ion Conductances, Dissociation Constants, and Speciation to 673 K and 28 MPa. The Journal of Physical Chemistry B 2005, 109, 9034-9046.

90. Kalinichev, A. G., Molecular Simulations of Liquid and Supercritical Water: Thermodynamics, Structure, and Hydrogen Bonding. Reviews in Mineralogy and Geochemistry 2001, 42, 83-129.

91. Alejandre, J.; Tildesley, D. J.; Chapela, G. A., Molecular Dynamics Simulation of the Orthobaric Densities and Surface Tension of Water. The Journal of Chemical Physics 1995, 102, 4574-4583.

92. Driesner, T.; Seward, T. M.; Tironi, I. G., Molecular Dynamics Simulation Study of Ionic Hydration and Ion Association in Dilute and 1 Molal Aqueous Sodium Chloride Solutions from Ambient to Supercritical Conditions. Geochimica Et Cosmochimica Acta 1998, 62, 3095-3107.

93. Reagan, M. T.; Harris, J. G.; Tester, J. W., Molecular Simulations of Dense Hydrothermal NaCl−H2O Solutions from Subcritical to Supercritical Conditions. The Journal of Physical Chemistry B 1999, 103, 7935-7941.

Page 172: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

152

94. Chialvo, A. A.; Simonson, J. M., Aqueous Na+Cl− Pair Association from Liquidlike to Steamlike Densities along Near-Critical Isotherms. The Journal of Chemical Physics 2003, 118, 7921-7929.

95. Kirkwood, J. G.; Buff, F. P., The Statistical Mechanical Theory of Solutions. I. The Journal of Chemical Physics 1951, 19, 774-777.

96. Newman, K. E., Kirkwood-Buff solution theory: derivation and applications. Chemical Society Reviews 1994, 23, 31-40.

97. Abdulagatov, I. M.; Dvoryanchikov, V. I.; Mursalov, B. A.; Kamalov, A. N., Measurements of the Heat Capacity at Constant Colume of H2O+Na2SO4 in Near-critical and Supercritical Water. Fluid Phase Equilibria 1998, 150–151, 525-535.

98. Kravchuk, K. G.; Todheide, K., Solid-liquid-vapour Equilibria in Water-Salt Systems: A DTA Study at Elevated Temperatures and Pressures. Z. Phys. Chemie-Int. J. Res. Phys. Chem. Chem. Phys. 1996, 193, 139-150.

99. Rouff, A. A.; Rabe, S.; Vogel, F., Phase Transitions in Hydrothermal K2HPO4 Solutions. The Journal of Supercritical Fluids 2011, 57, 207-212.

100. Benrath, A., Über die Löslichkeit von Salzen und Salzgemischen in Wasser bei Temperaturen oberhalb von 100°C. III. Zeitschrift für Anorganische Chemie 1941, 247, 147-160.

101. Benrath, A.; Gjedebo, F.; Schiffers, B.; Wunderlich, H., Über die Löslichkeit von Salzen und Salzgemischen in Wasser bei Temperaturen oberhalb von 100°C I. Zeitschrift für Anorganische Chemie 1937, 231, 285-297.

102. Booth, H. S.; Bidwell, R. M., Solubilities of Salts in Water at High Temperatures. Journal of the American Chemical Society 1950, 72, 2567-2575.

103. Keevil, N. B., Vapor Pressures of Aqueous Solutions at High Temperatures. J. Amer. Chem. Soc. 1942, 64, 841-850.

104. Khaibullin, I. K.; Novikov, B., Experimental Determination of the Parameters of Saturation of Solutions at High Temperatures by the Gamma-Ray Method. High Temp. 1973, 11, 276-282.

105. Marshall, W. L.; Wright, H. W.; Secoy, C. H., A Phase-Study Spparatus for Semimicro Use above Atmospheric Pressure. Journal of Chemical Education 1954, 31, 34.

106. Schroeder, W. C.; Gabriel, A.; Partridge, E. P., Solubility Equilibria of Sodium Sulfate at Temperatures of 150 to 350°C. I. Effect of Sodium Hydroxide and Sodium Chloride. Journal of the American Chemical Society 1935, 57, 1539-1546.

107. Schroeder, W. C.; Berk, A. A.; Partridge, E. P., Solubility Equilibria of Na2SO4 at Temperatures from 150 to 350°C. IV. Comparison of Evaporation and Equilibrium Solubility Values. Journal of the American Chemical Society 1937, 59, 1790-1795.

Page 173: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

153

108. Smits, A.; Wuite, J., On the System Water-Natrium Sulfate. Proc. R. Soc. Amsterdam 1909, 12, 244-257.

109. Tilden, W.; Shenstone, W., On the Solubility of Salts in Water at High Temperatures. Proc. R. Soc. London, Ser. A 1883, 35, 345-346.

110. Valyashko, V. M.; Abdulagatov, I. M.; Levelt Sengers, J. M. H., Vapor−Liquid−Solid Phase Transitions in Aqueous Sodium Sulfate and Sodium Carbonate from Heat Capacity Measurements near the First Critical End Point. 2. Phase Boundaries. Journal of Chemical & Engineering Data 2000, 45, 1139-1149.

111. Keevil, N. B., Vapor Pressures of Aqueous Solutions at High Temperatures. Journal of the American Chemical Society 1942, 64, 841-850.

112. Puchkov, L. V.; Kurochkina, V. V.; Matveeva, R. P., Deposit. VINITI USSR 1976, n.3474-76, p.20.

113. Ravich, M. I.; Borovaya, F.; Ketkovich, V. Y., Solubility and Vapor Pressure of Saturated Solutions in the System KCl - K2SO4 - H2O at High Temperatures. Izvestiya Sektora Fiziko-Khimicheskogo Analiza 1953, 22, 225-239.

114. Ravich, M. I.; Borovaya, F. E., Phase Equilibria in the System K2SO4 - H2O at Elevated Temperatures and Pressures. Zh. Neorg. Khim. 1968, 13, 1418-1425.

115. Marshall, W. L.; Hall, C. E.; Mesmer, R. E., The System Dipotassium Hydrogen Phosphate-Water at High Temperatures (100-400°C); Liquid-Liquid Immiscibility and Concentrated Solutions. J. Inorg. Nucl. Chem. 1981, 43, 722-728.

116. Marshall, W. L., Two-Liquid-Phase Boundaries and Critical Phenomena at 275-400°C for High-Temperature Aqeous Potassium Phosphate and Sodium Phosphate Solutions. Potential Applications for Steam Generators. J. Chem. Eng. Data 1982, 27, 175-180.

117. Reimer, J.; Vogel, F., High Pressure Differential Scanning Calorimetry of the Hydrothermal Salt Solutions K2SO4-Na2SO4-H2O and K2HPO4-H2O. RSC Advances 2013, 3, 24503-24508.

118. Ravich, M. I.; Valyashko, V. M., Compositions and Vapor Pressures of Eutonic Solutions in the System K2SO4 - KLiSO4 - H2O at Elevated Temperatures. Zhurnal Neorg. Khimii 1969, 14, 1650-1654.

119. Freyer, D.; Voigt, W., The Measurement of Sulfate Mineral Solubilities in the Na-K-Ca-Cl-SO4-H2O System at Temperatures of 100, 150 and 200°C. Geochimica Et Cosmochimica Acta 2004, 68, 307-318.

120. Schroeder, W. C.; Berk, A. A.; Gabriel, A., Solubility Equilibria of Sodium Sulfate at Temperatures from 150 to 350°C. II. Effect of Sodium Hydroxide and Sodium Carbonate. . Journal of the American Chemical Society 1936, 58, 843-849.

121. Blasdale, W. C.; Robson, H. L., The System Water and the Sulphates of Sodium and Magnesium. Journal of the American Chemical Society 1928, 50, 35-46.

Page 174: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

154

122. Marshall, W. L., Two-liquid-phase Boundaries and Critical Phenomena at 275-400°C; for High-Temperature Aqueous Potassium Phosphate and Sodium Phosphate Solutions. Potential Applications for Steam Generators. Journal of Chemical & Engineering Data 1982, 27, 175-180.

123. Gude, M.; Teja, A. S., Vapor-Liquid Critical Properties of Elements and Compounds. 4. Aliphatic Alkanols. J. Chem. Eng. Data 1995, 40, 1025-1036.

124. Peng, G. Methane Production from Microalgae via Continuous Catalytic Supercritical Water Gasification: Development of Catalysts and Sulfur Removal Techniques. PhD Thesis, EPFL, Lausanne, 2015.

125. Robson, H. L., The System MgSO4-H2O from 68 to 240°C. Journal of the American Chemical Society 1927, 49, 2772-2783.

126. Smits, A.; Rinse, J.; Louwe Kooymans, L. H., Systeme mit zurücklaufenden Schmelzlinien. II.Das System Wasser - Magnesiumsulfat. Zeitschrift für Physikalische Chemie 1928, 135, 78-84.

127. Akhumov, E. I.; Pilkova, E. V., Solubility and Supersaturation in the System Na2SO4 - H2O at High Temperatures. Zhurnal Neorg. Khimii 1958, 3, 2178-2183.

128. Baierlein, H. Zur Löslichkeit von Salzen in überkritischem Wasserdampf. PhD Thesis, Universität Erlangen-Nürnberg, Erlangen, 1983.

129. Reimer, J.; Steele-MacInnis, M.; Wambach, J. M.; Vogel, F., Ion Association in Hydrothermal Sodium Sulfate Solutions Studied by Modulated FT-IR-Raman Spectroscopy and Molecular Dynamics. J. Phys. Chem. B 2015, 119, 9847-9857.

130. Hazlebeck, D. A. Hydrothermal Processing with Phosphate Additive. US 6,238,568 B1, 1999.

131. Gorman, M. The Phase Behavior of Hydrothermal Salt Solutions; Paul Scherrer Institut: 29.8.2013, 2013.

132. Fulton, J. L.; Chen, Y.; Heald, S. M.; Balasubramanian, M., High-Pressure, High-Temperature X-Ray Absorption Fine Structure Transmission Cell for the Study of Aqueous Ions with Low Absorption-Edge Energies. Review of Scientific Instruments 2004, 75, 5228-5231.

133. Wernet, P.; Testemale, D.; Hazemann, J.-L.; Argoud, R.; Glatzel, P.; Pettersson, L. G. M.; Nilsson, A.; Bergmann, U., Spectroscopic Characterization of Microscopic Hydrogen-Bonding Disparities in Supercritical Water. The Journal of Chemical Physics 2005, 123, 154503.

134. Zhang, H.; Wang, S.; Sun, C. C., Ab Initio Investigation on Ion-associated Species and Association Process in Aqueous Na2SO4 and Na2SO4/MgSO4 Solutions. Journal of Chemical Physics 2011, 135, 084309.

Page 175: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

155

135. Watanabe, D.; Hamaguchi, H.-o., Ion Association Dynamics in Aqueous Solutions of Sulfate Salts as Studied by Raman Band Shape Analysis. The Journal of Chemical Physics 2005, 123, 034508.

136. Choi, B.-K.; Lockwood, D. J., Raman Spectrum of Na2SO4 (Phases I and II). Solid State Communications 1990, 76, 863-866.

137. Elenevskaya, V. M.; Ravich, M. I., Solubility of Li2SO4 and Li2CO3 at high temperatures. Zh. Neorg. Khim. 1961, 6, 2380-2386.

138. Cazzanelli, E.; Frech, R., Temperature Dependent Raman Spectra of Monoclinic and

Cubic Li2SO. J. Chem. Phys. 1984, 81, 4729.

139. Campbell, A. N., The System Li2SO4-H2O. Journal of the American Chemical Society 1943, 65, 2268-2271.

140. Waldeck, W. F.; Lynn, G.; Hill, A. E., Aqueous Solubility of Salts at High Temperatures. I. Solubility of Sodium Carbonate from 50 to 348°C. Journal of the American Chemical Society 1932, 54, 928-936.

141. Rudolph, W. W.; Irmer, G., Raman and Infrared Spectroscopic Investigations on Aqueous Alkali Metal Phosphate Solutions and Density Functional Theory Calculations of Phosphate-Water Clusters. Applied Spectroscopy 2007, 61, 1312-1327.

142. Marshall, W. L.; Begun, G. M., Raman-Spectroscopy of Aqueous Phosphate Solutions at Temperatures up to 450°C - 2 Liquid Phases, Supercritical Fluids and Pyro-Phosphate to Ortho-Phosphate Conversions. Journal of the Chemical Society-Faraday Transactions II 1989, 85, 1963-1978.

143. Marcus, Y., Molecular Properties of SCW. In Supercritical Water. A Green Solvent: Properties and Uses, Marcus, Y., Ed. John Wiley & Sons, Inc.: Hoboken, New Jersey, 2012.

144. Christen, M.; Hünenberger, P. H.; Bakowies, D.; Baron, R.; Bürgi, R.; Geerke, D. P.; Heinz, T. N.; Kastenholz, M. A.; Kräutler, V.; Oostenbrink, C.; Peter, C.; Trzesniak, D.; van Gunsteren, W. F., The GROMOS Software for Biomolecular Simulation: GROMOS05. Journal of Computational Chemistry 2005, 26, 1719-1751.

145. Scott, W. R. P.; Hünenberger, P. H.; Tironi, I. G.; Mark, A. E.; Billeter, S. R.; Fennen, J.; Torda, A. E.; Huber, T.; Krüger, P.; van Gunsteren, W. F., The GROMOS Biomolecular Simulation Program Package. The Journal of Physical Chemistry A 1999, 103, 3596-3607.

146. Van Gunsteren, W. F.; Billeter, S. R.; Eising, A. A.; Hünenberger, P. H.; Krüger, P.; Mark, A. E.; Scott, W. R. P.; Tironi, I. G., Biomolecular Simulation: The GROMOS96 Manual and User Guide. vdf Hochschulverlag AG an der ETH Zürich: 1996.

147. Van Gunsteren, W. F. www.gromos.net (accessed 01.03.2015).

148. Reif, M. M.; Hünenberger, P. H.; Oostenbrink, C., New Interaction Parameters for Charged Amino Acid Side Chains in the GROMOS Force Field. J. Chem. Theory Comput. 2012, 8, 3705-3723.

Page 176: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

156

149. Reif, M. M.; Hünenberger, P. H., Computation of Methodology-Independent Single-Ion Solvation Properties from Molecular Simulations. IV. Optimized Lennard-Jones Interaction Parameter Sets for the Alkali and Halide Ions in Water. The Journal of Chemical Physics 2011, 134, 144104.

150. Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P., The Missing Term in Effective Pair Potentials. The Journal of Physical Chemistry 1987, 91, 6269-6271.

151. Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R., Molecular Dynamics with Coupling to an External Bath. The Journal of Chemical Physics 1984, 81, 3684-3690.

152. Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. C., Numerical Integration of the Cartesian Equations of Motion of a System with Constraints: Molecular Dynamics of n-Alkanes. Journal of Computational Physics 1977, 23, 327-341.

153. Wernersson, E.; Jungwirth, P., Effect of Water Polarizability on the Properties of Solutions of Polyvalent Ions: Simulations of Aqueous Sodium Sulfate with Different Force Fields. J. Chem. Theory Comput. 2010, 6, 3233-3240.

154. Cannon, W. R.; Pettitt, B. M.; McCammon, J. A., Sulfate Anion in Water: Model Structural, Thermodynamic, and Dynamic Properties. The Journal of Physical Chemistry 1994, 98, 6225-6230.

155. Humphrey, W.; Dalke, A.; Schulten, K., VMD - Visual Molecular Dynamics. J. Molec. Graphics 1996, 14, 33-38.

156. Nord, A. G., Refinement of the Crystal Structure of Thenardite, Na2SO4 (V). Acta Chemica Scandinavica 1973, 27, 814-822.

157. Naruse, H.; Tanaka, K.; Morikawa, H.; Marumo, F.; Mehrotra, B. N., Structure of Na2SO4(I) at 693 K. Acta Crystallographica Section B 1987, 43, 143-146.

158. Ravich, M. I.; Borovaya, F. E., Phase Equilibria in the System Na2SO4 - H2O at High Temperatures and Pressures. Zh. Neorg. Khim. 1964, 9, 952-974.

159. Singh, A.; Puri, S., Phase Separation in Ternary Fluid Mixtures: A Molecular Dynamics Study. Soft Matter 2015, 11, 2213-2219.

160. Mysen, B.; Richet, P., Silicate Glasses and Melts, 1st Edition. Elsevier B.V.: Amsterdam, 2005.

161. Holmes, H. F.; Mesmer, R. E., Isopiestic Studies of NaHSO4 (aq) at Elevated Temperatures. Thermodynamic Properties. J. Chem. Thermod. 1993, 25, 99-110.

162. Kuhne, T. D., Second Generation Car-Parrinello Molecular Dynamics. Wiley Interdisciplinary Reviews-Computational Molecular Science 2014, 4, 391-406.

Page 177: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

References

157

163. Schubert, M.; Regler, J. W.; Vogel, F., Continuous salt precipitation and separation from supercritical water. Part 2. Type 2 salts and mixtures of two salts. Journal of Supercritical Fluids 2010, 52, 113-124.

164. Schubert, M.; Regler, J. W.; Vogel, F., Continuous salt precipitation and separation from supercritical water. Part 1: Type 1 salts. Journal of Supercritical Fluids 2010, 52, 99-112.

165. Schubert, M.; Mueller, J. B.; Vogel, F., Continuous Hydrothermal Gasification of Glycerol Mixtures: Effect of Glycerol and its Degradation Products on the Continuous Salt Separation and the Enhancing Effect of K3PO4 on the Glycerol Degradation. Journal of Supercritical Fluids 2014, 95, 364-372.

166. Schubert, M.; Mueller, J. B.; Vogel, F., Continuous Hydrothermal Gasification of Glycerol Mixtures: Autothermal Operation, Simultaneous Salt Recovery, and the Effect of K3PO4 on the Catalytic Gasification. Industrial & Engineering Chemistry Research 2014, 53, 8404-8415.

167. Müller, J. B.; Vogel, F., Tar and Coke Formation During Hydrothermal Processing of Glycerol and Glucose. Influence of Temperature, Residence Time and Feed Concentration. The Journal of Supercritical Fluids 2012, 70, 126-136.

168. Müller, J. B. Hydrothermal Gasification of Biomass - Investigation on Coke Formation and Continuous Salt Separation with Pure Substrates and Real Biomass. PhD Thesis, ETH Zurich, Zurich, 2012.

169. Wasserman, E.; Wood, B.; Brodholt, J., The Static Dielectric Constant of Water at Pressures up to 20 kbar and Temperatures to 1273 K: Experiment, Simulations, and Empirical Equations. Geochimica et Cosmochimica Acta 1995, 59, 1-6.

170. Jefferson, W. T.; Holgate, H. R.; Fred, J. A.; Paul, A. W.; William, R. K.; Glenn, T. H.; Herbert, E. B., Supercritical Water Oxidation Technology. In Emerging Technologies in Hazardous Waste Management III, American Chemical Society: 1993; Vol. 518, pp 35-76.

171. Viereck, S. PhD Thesis, ETH Zurich, Zurich, in preparation.

172. Ishikawa, D.; Cai, Y. Q.; Shaw, D. M.; Tse, J. S.; Hiraoka, N.; Baron, A. Q. R., X-ray Raman Scattering of Water Near the Critical Point: Comparison of an Isotherm and Isochore. arXiv:1210.4274v1 [physics.chem-ph] 2012.

173. Sahle, C. J.; Sternemann, C.; Schmidt, C.; Lehtola, S.; Jahn, S.; Simonelli, L.; Huotari, S.; Hakala, M.; Pylkkänen, T.; Nyrow, A.; Mende, K.; Tolan, M.; Hämäläinen, K.; Wilke, M., Microscopic Structure of Water at Elevated Pressures and Temperatures. PNAS 2013, 110, 6301-6306.

174. Soper, A. K., Computer Simulation as a Tool for the Interpretation of Total Scattering Data from Glasses and Liquids. Molecular Simulations 2012, 38, 1171-1185.

Page 178: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

List of Publications

158

Curriculum Vitae

Name: Joachim Reimer

Place of birth: Worms, Germany

Date of birth: January 19th, 1987

Nationality: German

EDUCATION

2012 – present Doctoral studies at the Institute for Chemical and Bioengineering, ETH

Zürich, Switzerland and Paul Scherrer Institute, Villigen, Switzerland

PhD thesis: Biomass Related Salt Solutions at Hydrothermal Conditions

Supervisors: Prof. Dr. A.Wokaun, Prof. Dr. F. Vogel

Co-examiner: Prof. Dr. M. Mazzotti

2011 Karlsruher Institut für Technologie (KIT), Germany

Diploma thesis: Untersuchungen zur heterogenkatalytischen

Hydrodesoxygenierung von Furfural und 5-Hydoxymethylfurfural

Supervisor: Prof. Dr. J.-D. Grunwaldt

2007-2011 Karlsruher Institut für Technologie (KIT), Germany

Diploma studies in Chemistry (specialization in Inorganic Chemistry)

1997-2006 Rudi-Stephan-Gymnasium, Worms, Germany

High school diploma

Page 179: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

List of Publications

159

Publications Related to this Thesis

Peer-reviewed publications (as of October 2015)

J. Reimer, F. Vogel, High Pressure Differential Scanning Calorimetry of the

Hydrothermal Salt Solutions K2SO4–Na2SO4–H2O and K2HPO4–H2O, RSC Adv., 2013,3,

24503-24508.

J. Reimer, M. Steele-MacInnis, J.M. Wambach, F. Vogel, Ion Association in

Hydrothermal Sulfate Solutions, studied by Modulated FT-IR-Raman and Molecular

Dynamics, J. Phys. Chem. B, 2015, 19 (30), 9847–9857.

M. Steele-MacInnis, J. Reimer, S. Bachmann, Hydrothermal properties of the COS/D2

water model: A polarizable charge-on-spring water model, at elevated temperatures

and pressures, RCS Adv., 2015, 5, 75846-75856.

J. Reimer, F. Vogel, Influence of Anions and Cations on the Phase Behavior of Ternary

Salt Solutions Studied by High Pressure Differential Scanning Calorimetry,

J. Supercritical Fluids, 2016, 109, 141-147.

Publications in preparation

J. Reimer, M. Steele-MacInnis, F. Vogel, Speciation and Structural Properties of

Hydrothermal Na-K-SO4 Solutions Studied by Molecular Dynamics Simulations, in

preparation.

J. Reimer, G. Peng, E. De Boni, S. Viereck, J. Breinl, F. Vogel, A Novel Salt Separator for

the Hydrothermal Gasification of Biomass, in preparation.

Patent applications

E. De Boni, J. Reimer, G. Peng, H. Zöhrer, F. Vogel, Paul Scherrer Institut, Patent

EP2918549 A1

Page 180: In Copyright - Non-Commercial Use Permitted Rights ...48490/eth-48490-02.pdf2-ratios of 1:10. Those results indicate that addition of K 2 HPO 4 can be used to help convert otherwise

List of Publications

160

Posters

J. Reimer, S. Pin, F. Vogel, Biomass-related Salt Solutions at Hydrothermal Conditions

- Investigation with High Pressure Differential Scanning Calorimetry, Summer School

of Calorimetry, CNRS, Lyon, France, poster session, June 10-15, 2012.

S. Pin, T. Huthwelker, J. Reimer, J.M. Wambach, F. Vogel, New high temperature -

high pressure XAS cell to study salt precipitation at and near supercritical conditions

of water, 3rd Workshop on the simultaneous combination of spectroscopies with x-

ray absorption, scattering and diffraction techniques, ETH Zürich, Switzerland, poster

session, July 4-6, 2012.

J. Reimer, S. Pin, T. Huthwelker, F. Vogel, New high temperature - high pressure XAS

cell to study salt precipitation at and near supercritical conditions of water, 3rd Joint

Users’ Meeting at PSI: JUM@P’, Paul Scherrer Institute, Villigen, Switzerland,

September 18-20, 2013.

J. Reimer, F. Vogel, Biomass-related Salt Solutions at Hydrothermal conditions,

Biomass for Swiss Energy Future Conference, Paul Scherrer Institute, Villigen,

Switzerland, December 2, 2014.

Conference talks

J. Reimer, S. Pin, T. Huthwelker, F. Vogel, Biomass related Salt Solutions at

Hydrothermal Conditions: Investigations with Isochoric Differential Scanning

Calorimetry, 16th International Conference on the Properties of Water and Steam

ICPWS, London, UK, September 1-5, 2013.

J. Reimer, J.M. Wambach, F. Vogel, Modulated FT-IR-Raman Spectroscopy of

Hydrothermal Salt Solutions, 11th International GeoRaman Conference, St. Louis,

Missouri, USA, June 15-19, 2014.