<<

View Article Online View Journal PCCP Physical Chemistry Chemical Physics Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: S. Tischer, M. Börnhorst, J. Amsler, G. Schoch and O. Deutschmann, Phys. Chem. Chem. Phys., 2019, DOI: 10.1039/C9CP01529A.

Volume 19 Number 1 This is an Accepted Manuscript, which has been through the 7 January 2017 Pages 1-896 Royal Society of Chemistry peer review process and has been PCCP accepted for publication. Physical Chemistry Chemical Physics Accepted Manuscripts are published online shortly after acceptance, rsc.li/pccp before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the Information for Authors.

Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard

ISSN 1463-9076 Terms & Conditions and the Ethical guidelines still apply. In no event

PAPER H.-P. Loock et al. shall the Royal Society of Chemistry be held responsible for any errors Determination of the thermal, oxidative and photochemical degradation rates of scintillator liquid by fluorescence EEM spectroscopy or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains.

rsc.li/pccp Page 1 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A

Thermodynamics and reaction mechanism of de- composition †

Steffen Tischer,∗a Marion Börnhorst,b Jonas Amsler,b Günter Schoch,b and Olaf Deutschmanna,b Manuscript The selective catalytic reduction technique for automotive applications depends on pro- duction from a urea- solution by thermolysis and hydrolysis. In this process, undesired liquid and solid by-products are formed in the exhaust pipe. The formation and decomposition of these by-products have been studied by thermogravimetric analysis and differential scanning calorime- try. A new reaction scheme is proposed that emphasizes the role of thermodynamic equilibrium of the reactants in liquid and solid phases. Thermodynamic data for have been refined. The observed phenomenon of liquefaction and re-solidification of in the temperature range ◦

193–230 C is explained by formation of a eutectic mixture with urea. Accepted

Creative Commons Attribution-NonCommercial 3.0 Unported Licence. 1 Introduction from experimental results and literature data, 23 possible reac- Air pollution by nitrogen oxides from Diesel engines is a main tions including urea and its by-products biuret, , am- problem concerning environment and society. Therefore, gov- melide, and are presented. Further, cyanate ernments follow the need to regulate emissions by law (e.g. and cyanurate salts and cyanamide are proposed as possible in- 715/2007/EG, "Euro 5 and Euro 6"). 1 The favored method to termediates of high temperature urea decomposition. Triuret pro- Physics reduce nitrogen oxides is selective catalytic reduction (SCR) us- duction and decomposition is not accounted for in this study. ing ammonia as reducing agent. Due to security issues, passenger The authors classify urea decomposition into four temperature cars cannot be equipped with an ammonia tank. Consequently, regions. The first temperature regime from room temperature This article is licensed under a ammonia is provided in form of a urea-water solution, which de- to 190 ◦C comprises urea melting and vaporization starting from composes through thermolysis and hydrolysis when dispersively 133 ◦C. With increasing temperature, urea decomposes to am- dosed into the hot exhaust pipe. During the desired decomposi- monia and isocyanic acid, the latter leading to biuret, cyanuric

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. ◦ tion of urea at temperatures above 130 C undesired intermedi- acid and formation. The second temperature region ◦ ates and by-products in liquid and solid form occur and stick to of 190–250 C is dedicated to biuret decomposition accompanied Chemical the wall of the exhaust pipe due to the inevitable spray-wall in- by several side reactions forming cyanuric acid and ammelide. At teraction. 2–4 These condensed deposits in the exhaust pipe bear 225 ◦C the melt is observed to be converted into a sticky, solid the risk of increased pressure drop and insufficient conversion matrix, which is assumed to originate from ionic formations of to ammonia. Thus there is a need to understand the decompo- different by-products without evidence. Besides small amounts sition kinetics of urea and its by-products, that has been exten- of ammelide, ammeline and melamine, cyanuric acid is the main sively studied by many authors. 5–11 Common experimental meth- component observed at 250 ◦C. The third temperature range from ods include Thermogravimetric Analysis (TG), Differential Scan- 250 to 360 ◦C represents the sublimation and decomposition of ning Calorimetry (DSC), High Performance Liquid Chromatogra- cyanuric acid. Ammelide, ammeline and melamine are proposed Chemistry phy (HPLC) and Fourier transform-infrared spectroscopy (FT-IR). to gradually decompose at temperatures above 360 ◦C marking A first detailed description of urea decomposition behavior was the fourth temperature region. The authors state elimination of proposed by Schaber et al. 8 based on TG, HPLC, FT-IR and am- ammelide at 600 ◦C and ammeline at 700 ◦C. High temperature monium ISE (ion-selective electrode) measurements. Concluding residues are not investigated. 8

Eichelbaum et al. 10 propose a reaction network for urea de- a Institute of Catalysis Research and Technology (IKFT), Karlsruhe Institute of Technol- composition consisting of nine major reactions based on simul- ogy (KIT), Karlsruhe, Germany. Fax: +49 721 608 44805; Tel: +49 721 608 42114; taneous TG/DTA (Differential Thermal Analysis) coupled with Physical E-mail: [email protected] GC/MS (Gas Chromatography/Mass Spectrometry) gas analyses. b Institute for Chemical Technology and Polymer Chemistry (ITCP), Karlsruhe Institute Decomposition reactions of ammelide, ammeline and melamine of Technology (KIT), Karlsruhe, Germany. † Electronic Supplementary Information (ESI) available: [details of any supplemen- are defined and total decomposition is observed for temperatures ◦ tary information available should be included here]. See DOI: 00.0000/00000000. above 625 C. However, the proposed reaction scheme lacks in

1–13 | 1 Physical Chemistry Chemical Physics Page 2 of 14 View Article Online DOI: 10.1039/C9CP01529A

description of relevant parallel and equilibrium reactions of urea Table 1 Experiments used for this study by-products. Acceleration of urea pyrolysis by different metal 10 Type Crucible Substance Initial Ramp exchanged zeolites is demonstrated. A more detailed reaction weight (mg) (K/min) scheme derived by flow reactor experiments using FTIR spec- TG plate urea 6.23 2 troscopy for gaseous and HPLC for solid reaction product analysis TG cylinder urea 6.18 2 11 TG cylinder urea 60.3 2 covers 15 decomposition reactions. For the first time, triuret TG cylinder urea 10.8 10 production and decomposition is included and several reactions TG cylinder 32.5 wt-% UWS 27.5 10 are proposed to be equilibrium reactions. TG plate biuret 5.24 2 TG cylinder biuret 5.18 2 Based on the proposed reaction schemes, a first kinetic model TG cylinder biuret 9.8 10 TG cylinder biuret 95.8 2 (to 195 ◦C) for evaporation and decomposition of urea water solution was de- ◦ 12 TG cylinder biuret 94.6 2 (to 200 C) veloped by Ebrahimian et al. The model describes urea decom- TG cylinder biuret 94.4 2 (to 210 ◦C) position to ammonia and isocyanic acid and the equilibrium re- TG cylinder triuret 6.98 2 action forming biuret. Reactions from biuret to cyanuric acid and TG cylinder triuret 10.31 10 TG plate cyanuric acid 5.87 2 from cyanuric acid to ammelide and isocyanic acid are included. TG cylinder cyanuric acid 5.37 2 Manuscript Ammelide is assumed to decompose to gaseous by-products. TG cylinder cyanuric acid 10.1 10 Current kinetic models are based on the reaction network pro- TG plate ammelide 5.58 2 TG cylinder ammelide 8.72 2 11 posed by Bernhard et al. and validated against TG and HPLC TG cylinder ammelide 9.34 10 experimental data. The model includes formation and decompo- DSC cylinder urea 14.2 2 sition reactions of the most relevant by-products and reproduces DSC cylinder biuret 12.8 2 DSC cylinder triuret 5.9 2 the characteristic decomposition stages of urea adequately. 13 DSC cylinder cyanuric acid 16.4 2 However, important physical and chemical processes are not ac-

counted for. A biuret matrix species is defined to cover the effect Accepted ◦ of solidification at 220 C. Ammelide decomposition is modeled gravimetrical analysis, a Netzsch STA 409 C was equipped with Creative Commons Attribution-NonCommercial 3.0 Unported Licence. as sublimation while further high molecular by-products are not the thermal controller TASC 414/2. The following standard pro- included in the model. Further, recent investigations have shown cedure is conducted for each deposit sample and for reference that the current kinetic model lacks in reproduction of mass trans- measurements. Representative samples are ground and placed in fer effects at the sample surface and in the prediction of reactions a corundum crucible with an initial sample mass of 5–100 mg. including gas phase species. The samples are heated from 40 to 700 ◦C at a constant heating Physics Thermogravimetric measurements have shown a strong influ- rate of 2 or 10 K/min. TG analysis is performed in synthetic air ence of the experimental boundary conditions on urea decompo- (20.5% O2 in N2) using a purge gas flow rate of 100 mL/min. sition kinetics. Besides the sample heating rate 5,6,8, decomposi- Different geometries of corundum crucibles are used to hold the This article is licensed under a tion behavior is highly sensitive to the geometric arrangement of samples during measurement: a cylinder-type crucible with an the samples and respective crucibles. 6,10 Increased surface area inner diameter of 6 mm and a height of 12 mm and a plate-type of the sample is assumed to accelerate mass transport of gaseous crucible of 15 mm inner diameter and a height of 5 mm. Cylinder Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. products. 10,11 Further, the presence of water is stated to decrease geometry, sample mass and heating rate are systematically varied 10,11 by-product formation due to isocyanic acid hydrolysis. for different samples in order to derive a large database for kinetic Chemical In a very recent publication by Wang et al. 14 the gaseous and modeling. Pure urea (Merck, ≥ 99.5%), biuret (Sigma Aldrich, condensed products of the decomposition of urea have been ana- ≥ 98%), triuret (Sigma Aldrich, ≥ 95%), cyanuric acid (Sigma lyzed in detail for the temperature range from 132.5 to 190 ◦C Aldrich, ≥ 98%), ammelide (Dr. Ehrenstorfer GmbH, 99%), am- by means of online mass spectrometry and liquid chromatog- meline (Sigma Aldrich, 97.9%) and a 32.5 wt-% urea water so- raphy mass spectrometry, respectively. They derived a reaction lution (UWS) are used for the measurements. TG analysis gives scheme where biuret and cyanuric acid are formed directly by information about the mass loss of a sample by evaporation and self-combination reactions of urea. reactions under specified conditions and therefore gives informa- So far, available kinetic schemes for chemical decomposition tion about the decomposition behavior. Chemistry are not suitable for accurate description of reactions in and be- DSC is a thermo-analytical measurement technique to deter- tween all phases. This is our motivation to systematically inves- mine the difference in heat required to increase the sample tem- tigate the decomposition of urea and its subsequent products bi- perature compared to a reference sample. A Mettler DSC 30 uret, triuret, cyanuric acid, and ammelide by TG and DSC exper- is used to measure the thermal properties of urea and its by- iments. Our numerical model is based on a consistent thermody- products. The calorimetry is operated under air flow applying a ◦ namic description of all phases. Using these data, a new kinetic heating rate of 2 K/min from 25 to 600 C. Initial sample weight scheme of urea decomposition is proposed. amounts to 10–15 mg. Aluminium crucibles are used as sam- ple holder and reference. All other experimental conditions were Physical 2 Experiments kept the same as in TG experiments. This allows correlation of Kinetic and thermodynamic data on the decomposition of urea measured thermal and kinetic properties and yields valuable data and its by-products are derived from thermogravimetrical (TG) for model development. analysis and differential scanning calorimetry (DSC). For thermo- A list of all experiments used in this study is given in Table 1.

2 | 1–13 Page 3 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A

⊖ 3 Numerical model pressed in terms of concentrations as ai = ci/ci with a reference ⊖ ⊖ ⊖ concentration c = p /RT for gas-phase species and c = ρi/Mi The TG and DSC experiments are simulated by a zero- i i for condensed species. Then the rate of a reverse reaction can be dimensional batch-type reactor model. It was implemented as linked to the equilibrium constant DETCHEMMPTR (MPTR = multi-phase tank reactor) into the ( ⊖ ) DETCHEM™ program package. 15 ∆ G K = exp − Rk , (8) p,k RT The reactor model consists of a set of species Si grouped into sets of phases Pj. Each species belongs to exactly one phase, i. e. ⊖ν K = K · ∏ c ik , (9) a phase transition of a chemical substance is handled by two dif- c,k p,k i Si∈Rk ferent species. Each species is associated with thermodynamic data in form of the NASA polynomials. 16 The molar heat capac- reverse rk rk = . (10) ity, molar enthalpy and molar entropy are computed as function Kc,k of temperature based on 7 coefficients a ...a . 1i 7i A special case of a heterogeneous process is the phase tran-

sition between liquid and gas. For the condensation, we may Manuscript cp,i 2 3 4 assume that molecules hitting the phase boundary stick with an = a1i + a2iT + a3iT + a4iT + a5iT (1) R accumulation factor αc. Thus, kinetic gas theory yields √ Hm,i a2i 2 a3i 3 a4i 4 a5i 5 RT = a1iT + T + T + T + T + a6i (2) condensation α gas R 2 3 4 5 rk = c ci . (11) 2πMi S a a a m,i = a ln(T) + a T + 3i T 2 + 4i T 3 + 5i T 4 + a (3) Applying the definition of the reverse rate (Eq. 10), we get the R 1i 2i 2 3 4 7i Herz-Knudsen equation for evaporation 17 √ Accepted The molar volume of a species is either defined by ideal gas-law liq evaporation RT c Creative Commons Attribution-NonCommercial 3.0 Unported Licence. for gaseous species V , = RT/p or by assuming a constant density α i m i rk = c (12) 2πMi h for condensed species Vm,i = ρi/Mi. The phases are separated. Each phase occupies a volume with the Henry constant

V = ∑ V . (4) ρliq j m,i i RT Si∈Pj h = vap . (13) pi Mi Physics The concentrations are expressed locally with respect to the The batch-type reactor model consists of conservation equa- corresponding phase, i. e. c = n /V . The reaction rates are mostly i i j tions for species and enthalpy:

This article is licensed under a given in terms of an extended Arrhenius expression (see also ap- pendix A in supplemental information). For reaction R we can dni k = ∑n˙ (14) dt ik write a molar rate Rk ( ) Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. β Ea,k ν ( ) k − ˜ik˜ dH rk = AkT exp ∏ ci˜ (5) = Ak T extern − T (15) RT ∈ W Chemical Si˜ Rk dt

Homogeneous reactions are reactions with reactants from the where temperature and total enthalpy are linked by same phase. However, the products of homogeneous reactions H = ∑ni · Hi(T) (16) may be released to a different phase. Let n˙ be the rate of pro- ik Si duction of species Si by reaction Rk. Thus, kW is a heat transfer coefficient. Since both, the TG and DSC n˙ik = Vjνikrk reactants ⊂ Pj. (6) experiments are driven by an external temperature profile, the value of the heat transfer coefficients is not very sensitive. It just

Heterogeneous reactions are assumed to occur at the interface has to be finite, otherwise the solution would jump in case of Chemistry of two phases. The reactants can come from both phases, but phase transitions. For the simulation reported here a value of can also come from only one of them. In our case, the contact −2 −1 kW = 200Wm K has been chosen. area between the phases shall be the cross-sectional area A of the The system of differential-algebraic equations is solved by the crucible, i. e. the phases are considered to be stacked on top of solver LIMEX. 18 The heat signal measured by DSC equals the each other in the cylindrical reactor. For an Arrhenius type of (negative) enthalpy change of the condensed phases in the cru- reaction, the production rate is likewise cible. Since gas-phase reactions shall not be considered in this

study, this enthalpy change equals the change of total enthalpy Physical n˙ik = Aνikrk. (7) minus the power to heat the gas.   All phases are considered to be ideal mixtures. Thus, the chem- ⊖ ical activity a of a species is p /p for the gas-phase and the DSC dH   dT i i P = − + ∑ nicp,i (17) dt gas dt molar fraction for all other phases. The activities can also be ex- Si

1–13 | 3 Physical Chemistry Chemical Physics Page 4 of 14 View Article Online DOI: 10.1039/C9CP01529A

4 Thermodynamic data ration mass fraction: − T Previous models 10,13 did not pay much attention to the thermo- Y (T) = 5.1687 · 10 3 · − 1.0108 . (24) urea(aq) K dynamics of the system. TG experiments can be simulated in batch-type reactor models by kinetic models without solving a There are no experimental data for higher temperatures. The sim- conservation equation for heat. With the availability of DSC data, ulation, however, requires a continuous polynomial throughout thermodynamic aspects of the mechanism can be studied in de- the entire interval (300–800 K). The extrapolation obeys two con- tail. This was the motivation for the authors to gather information straints: a non-negative heat capacity for the new species and an about the thermodynamic properties of all involved species. upper limit for the to assure a mole fraction of urea(aq)

For fundamental gaseous species, namely water H2O(g) and below unity. ammonia NH3(g), as well as for liquid water H2O(l) the standard Likewise, the chemical potentials of NH3(aq) and HNCO(aq) values from the thermodynamic database of DETCHEM 15 have were determined. In these cases the reference chemical potentials been used. For other species, values from literature were taken as are given by their respective gas-phase species. given in Table 2. NASA coefficients for these species were fitted µ0 = µ0 − RT (X ) Manuscript for a temperature range from 300 K to 800 K (see supplemental i(aq) i(g) ln i(aq) (25) material). The temperature dependency of solubility of ammonia was fitted to literature data 33–35 for temperature interval from 0 to 100 ◦C. Condensation of isocyanic acid ( ) ( ) 3 2 In the reaction mechanism by Brack et al. 13, isocyanic acid plays · −7 · T − · −5 · T YNH3(T) = 1.055 10 6.81 10 a major role. Even though isocyanic acid evaporates at 23.5 ◦C, K K reactions occurring at temperatures exceeding 133 ◦C were for- −3 T + 6.549 · 10 · + 1.61 (26) Accepted mulated in the condensed phase. For a thermodynamically con- K sistent formulation it is therefore necessary to look at the gas- Creative Commons Attribution-NonCommercial 3.0 Unported Licence. For isocyanic acid the temperature dependency of Henry’s co- liquid-equilibrium of isocyanic acid. Both are in equilibrium if efficient KH = cHNCO(aq)/pHNCO(g) was approximated for 25 to µ µ 100 ◦C by van’t Hoff’s extrapolation 36 HNCO(l) = HNCO(g) (18) ( ( )) ( ) ⊖ ⊖ ∆ H 1 1 pvap K (T) = K · exp sol − (27) µ0 µ0 HNCO H H ⊖ = + RT ln (19) R T T Physics HNCO(l) HNCO(g) p⊖ 37 ⊖ −1 −1 with a reference Henry coefficient KH = 26molL atm and where µ0 = H − TS is the chemical potential of the undi- ⊖ −1 i m,i m,i enthalpy of solution ∆solH = −34kJmol .

This article is licensed under a luted species. From the temperature dependency of the chemical potential of HNCO(l), we can derive the coefficients of the ther- Melting of urea and biuret modynamic polynomials by linear regression. The vapor pressure

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. follows an Antoine equation based on experimental data by Lin- hard et al. 28 ( ) Chemical vap p 1252.195 log HNCO = 4.69 − (20) 10 bar T/K − 29.167

Aqueous phase Urea, ammonia and isocyanic acid were also considered to be sol- uble in liquid water. In our model, these species form an ideally mixed aqueous phase (aq). Using literature data about the solu- bility, thermodynamic data were derived. At point of saturation, Chemistry the chemical potentials of a species must be the same in both phases, i. e. for urea:

µurea(aq) = µurea(s) (21)

µ0 µ0 urea(aq) + RT ln(Xurea(aq)) = urea(s) (22) Physical µ0 µ0 − Fig. 1 Gibbs free energy of urea and biuret urea(aq) = urea(s) RT ln(Xurea(aq)) (23)

Here, it is assumed that the mixture is ideal without interaction of With the literature data in Table 2 we can look at the phase the molecules (fugacity factor is unity). Experimental data 29–32 change from solid to liquid. For the pure substances the chemi- were fitted in a temperature range of 0–70 ◦C yielding the satu- cal potential equals the Gibbs free energy. The phase transition

4 | 1–13 Page 5 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A

Table 2 Thermodynamic properties of secondary products of urea decomposition in different phases (g=gas, l=liquid, s=solid)

⊖ ⊖ Species Symbol(Phase) ∆fH (kJ/mol) S (J/mol/K) cp (J/mol/K) Reference th 15,16 Water H2O(l) -285.828 69.939 4 order polynomial (75.351 at 298 K) th 15,16 H2O(g) -241.825 188.828 4 order polynomial (33.588 at 298 K) th 15,16 Ammonia NH3(g) -45.567 192.474 4 order polynomial (34.597 at 298 K) Urea urea(s) -333.599 105.9 93 19,20 urea(l) -319.7 140.15 93 21 urea(g) -235.55 282.94 26.84 + 0.2T − 1 · 10−4T 2 19 Biuret biu(s) -563.70 146.1 131.3 21,22 biu(l) -537.06 203.27 93 21,22 biu(g) -437.30 354.31 72.198 + 0.237T − 8.847 · 10−5T 2 − 1.455 · 106T −2 19 Triuret triu(s) -746.7 146.1 131.3 23 Cyanuric acid cya(s) -703.5 142.20 130.0 24,25 cya(g) -564.1 339.37 15.74 + 0.42T − 2 · 10−4T 2 24 Isocyanic acid HNCO(g) -101.7 238.229 15.34 + 0.126T − 9.523 · 10−5T 2 + 6.49 · 104T −2 (T ≤ 400K) 19 45.24 + 0.0307T − 7.619 · 10−6T 2 − 8.96 · 105T −2 (T ≥ 400K) Ammelide ammd(s) -492.9 149.10 94.958 + 0.262T − 8.444 · 10−5T 2 − 2.639 · 106T −2 26 Manuscript ammd(g) -303.9 429.87 15.74 + 0.42T − 2 · 10−4T 2 27 Ammeline ammn(s) -300.0 149.10 94.958 + 0.262T − 8.444 · 10−5T 2 − 2.639 · 106T −2 26

occurs when µ µ i(s) = i(l) . (28) Figure 1 plots the Gibbs free energy of urea and biuret in solid and Accepted liquid phases. As expected for urea, the two graphs intersect at ◦ Creative Commons Attribution-NonCommercial 3.0 Unported Licence. T = 133 C. However, in case of biuret the predicted melting tem- perature is 233 ◦C. This seems to be contradictory, because the common literature value 8,38 is between 190 and 193 ◦C. How- ever, it was also observed by Schaber 8 and Brack 13 that biuret starts decomposing at 193 ◦C when it turned into a liquid. The ◦ decomposition slowed down around 210 C and the melt formed Physics a foam-like structure. Between 220 and 230 ◦C the consistence was described as a matrix-like aggregate before a second decom- position step sets in. Analysis of the substance revealed that it still This article is licensed under a mainly consisted of biuret. Thus Brack introduced two phases of biuret, namely biu(melt) and biu(matrix) to model this behavior.

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. The authors want to give a thermodynamic explanation for this observation. It can be explained by an eutectic mixture of urea Fig. 2 Phase diagram of an eutectic mixture of urea and biuret Chemical and biuret. For varying mole fraction, we can plot the tempera- tures where µ0 µ0 products like triuret, cyanuric acid or ammelide. Around 230 ◦C, i(s) = i(l) + RT ln(Xi(l)) . (29) the biuret becomes liquid again and a second decomposition step Figure 2 shows the phase diagram of the binary urea/biuret sys- is observable in TG measurements. The third decomposition step tem. It agrees well with the results by Voskov et al. 39 This means ◦ between 330 and 400 C is associated with the sublimation of that biuret is a liquid at 193 ◦C if the liquid phase consists of 67% cyanuric acid. The remaining solid substances decompose at tem- biuret and 33% urea. As decomposition or evaporation of urea ◦ peratures above 400 C, for which we include a reaction path becomes faster at temperatures above 210 ◦C, biuret will become through ammelide, but this shall not be the focus of this paper. Chemistry solid again. As we will see later, there is another thermodynamic argument for the apparent melting point of biuret. 5 Reaction mechanism This interpretation yields a qualitative understanding of exper- Cyanuric acid imental DSC data for urea and biuret. Figure 3 shows a compar- In Brack’s model 13 cyanuric acid decomposes directly to isocyanic ison of the DCS signals. Together with TG data, we can identify acid (cya(s) 3 HNCO(g)) by a zero-th order reaction. In or- ◦ several processes. The sharp peak at 133 C indicates the melt- der to investigate this hypothesis, TG and DSC experiments were

ing of urea. The decomposition of urea starts more or less with run with pure cyanuric acid as initial sample. The TG experi- Physical the phase change. Biuret as a reaction product will remain in ment (Fig. 4) was conducted using two crucibles with different ◦ the liquid phase. At 193 C the decomposition process also starts diameters. If the reaction were a zero-th order reaction in the ◦ for biuret. The reaction slows down at around 210 C, when a homogeneous phase, the size of the crucible would not matter. foam-like structure is formed. Presumably, there is no liquid urea However, the reaction is faster in a flat crucible with larger sur- present anymore. Endothermic reactions may lead to follow-up face area than in a tall crucible. This suggests that the reaction

1–13 | 5 Physical Chemistry Chemical Physics Page 6 of 14 View Article Online DOI: 10.1039/C9CP01529A Manuscript

Fig. 3 Qualitative comparison of DCS signals for urea and biuret decom- Fig. 5 Comparison of DSC signal of decomposition of cyanuric acid position (1 - melting of urea, 2 - melting of biuret in an eutectic mixture, 3 (model 1: sublimation, model 2: direct decomposition to isocyanic acid) - urea disappears, 4 - melting of biuret, 5 - sublimation of cyanuric acid)

Triuret Accepted takes place at the surface of the sample. Indeed, both TG curves In the previous mechanism by Brack 13, triuret was a side prod- Creative Commons Attribution-NonCommercial 3.0 Unported Licence. can be modeled by a first-order surface reaction with an activa- uct of biuret reacting with isocyanic acid. It could have been tion energy of 141.3 kJ/mol. However, this immediately raises the omitted without significantly changing the predicted amounts of question about the resulting products, because the direct decom- cyanuric acid or ammelide. Looking at the chemical structures of position to isocyanic acid is an endothermic process consuming the latter two, there is good argument that triuret may decom- 352 kJ/mol. A comparison between experimental and expected pose to cyanuric acid by separation of ammonia and to ammelide

DSC signal (Fig. 5) shows that for a direct decomposition the DSC by separation of water. Thus, instead of going directly from bi- Physics signal should be larger by a factor of 2.5. The activation energy uret to cyanuric acid and ammelide by complex multi-molecule obviously agrees with the heat of sublimation of cyanuric acid. reactions, we may assume that triuret is an intermediate species

This article is licensed under a Therefore, it must be concluded that cyanuric acid evaporates di- of these processes. rectly into the gas-phase.

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. cya(s) cya(g) ∆H = 141.3kJ/mol (30) Chemical Chemistry

Fig. 6 Experimental TG data for decomposition of biuret and triuret

Figures 6 and 7 show a comparison of experimental TG and Physical DSC data for biuret and triuret. It can be noted that there are Fig. 4 Experimental TG data for decomposition of cyanuric acid in cylin- parallels in decomposition steps for both substances. A major de- der (diameter 6 mm) and plate-type (15 mm) crucibles composition step is observed for temperatures above 192 ◦C. The mass loss is mainly due to the release of isocyanic acid and am-

6 | 1–13 Page 7 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A Manuscript

Fig. 7 Experimental DSC data for decomposition of biuret and triuret Fig. 8 Fitting of thermodynamic data of triuret such that equilibrium cal- culation yield the same mass loss as TG experiments

Biuret monia. For biuret, as stated before, a second decomposition step Accepted appears at around 235 ◦C. This is not present for triuret. The pos- As seen for triuret, we cannot look at the decomposition of one Creative Commons Attribution-NonCommercial 3.0 Unported Licence. sible reactions to cyanuric acid and ammelide would both yield substance independently. One always has to keep in mind the a mass loss of about 12% and almost no heat signal of 8 and thermodynamical ensemble consisting of isocyanic acid, urea, bi- 22 kJ/mol, respectively. The decomposition step beyond 260 ◦C uret and triuret. In order to better understand the processes of ◦ can mainly be associated with the sublimation of cyanuric acid. biuret decomposition for temperatures between 193 and 210 C, A shift in temperature can be observed with respect to the de- we conducted TG experiments where biuret was heated up to

◦ Physics composition steps of both experiments. This is attributed to the 195, 200 and 210 C, respectively. Initially a constant temper- difference of the initial sample mass. ature ramp of 2 K/min was applied. After reaching the target temperature, it was kept constant. Figure 9 shows the result for Triuret is a highly unstable substance. The TG data (Fig. 6) ◦

This article is licensed under a ◦ the case 195 C. already show a mass loss for temperatures below 192 C. Since there is no significant DSC signal (besides the heat capacity of tri- uret) for these low temperatures, the mass loss might have been Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. due to kinetically limited reactions to isocyanic acid or ammelide,

which are almost isoenthalpic. However, all fits of kinetic param- Chemical eters over-predicted the produced amounts of these two prod- ucts for higher temperatures. Therefore it was concluded that the underlying process is a thermodynamically controlled equi- librium. Using the program DETCHEMEQUIL 15, equilibrium cal- culations for an ensemble consisting of urea(s), urea(l), biu(s), biu(l), triu(s), HNCO(l), HNCO(g) and N2 were carried out. Ac- cording to the literature data (Table 2), triuret would have re- mained stable up to 192 ◦C. However, literature data for triuret Chemistry are very insufficient. Thus, we tried to reverse-engineer thermo- dynamic data for triuret, by fitting the values of Gibbs free energy of triuret in such a way that the calculated equilibrium composi- tion shows the same mass loss (due to release of HNCO(g)) as the TG experiment (Fig. 6). The calculation becomes sensitive to the ◦ dilution of HNCO(g) by inert gas (e. g. N2). Figure 8 shows the Fig. 9 Experimental TG data for biuret that was first heated to 195 C results of the fitting for varying degrees of dilution. For the calcu- with a constant heating ramp and then temperature was kept constant lation of thermodynamic coefficients for triu(s), we have chosen a Physical dilution of 90%, i. e. the initial mixture consisted of 1 mol triuret For this sample, decomposition already starts below 193 ◦C. and 9 mol N2. As result we get standard enthalpy of formation Most notable feature of this graph is that its slope is larger for ⊖ ⊖ ∆fH = −708.9 kJ/mol and standard entropy S = 359.1 J/mol/K. the temperature ramp than for the case of constant temperature. The molar heat capacity was kept constant at cp = 131.3 J/mol/K. This indicates that the decomposition is not only kinetically con-

1–13 | 7 Physical Chemistry Chemical Physics Page 8 of 14 View Article Online DOI: 10.1039/C9CP01529A

trolled. Otherwise the reaction rate (and thus the slope of the graph) should be lower for lower temperatures. Therefore again, the mass loss during the temperature ramp is dominated by ther- modynamic equilibrium processes. A fact that is not well taken into account in experiment design is that this equilibrium depends on the dilution of HNCO(g) in the gas phase. Thus, the decom- position of biuret may start for temperatures as low as 170 ◦C for high purging rates with inert gas. On the other hand, if we calculate the equilibrium composition based on a lower dilution for an initial mixture of 1 mol biuret and 1 mol nitrogen using DETCHEMEQUIL 15, we get a surprisingly nice result (Figure 10). Here, we see a phase change at 192.5 ◦C when an eutectic mix- ture of liquid biuret and urea plus some triuret and gaseous iso-

cyanic acid becomes thermodynamically favored. This explains Manuscript the apparent melting point of biuret in literature.

Fig. 11 Comparison of TG and DSC data for the initiation of urea de- composition

1.87 J. However, the direct decomposition would have required Accepted 5.77 J. All numerical simulations containing a direct decomposi-

Creative Commons Attribution-NonCommercial 3.0 Unported Licence. tion reaction showed that TG and DSC signal could not be satis- fied at the same time. They differed by a factor of 3. Thus, we had to discard the direct decomposition reaction. Several ideas have been tried to initiate urea decomposition. Since the rate is sensitive to the size of the crucible, it was con-

cluded that the process occurs at the gas-liquid interface. The net Physics reaction that best fitted both, the DSC and the TG data in the interval 140–180 ◦C was

This article is licensed under a ∆ 2 urea(l) biu(l) + NH3(g) RH = 55.6kJ/mol (32)

Fig. 10 Equilibrium composition of an initial mixture of 1 mol biuret and as was also recently suggested by Wang et al. 14

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. 1 mol N2

Kinetic parameters Chemical

Urea Finally, we need to look at the initiation of urea decomposition, which starts with melting at 132 ◦C. Figure 11 shows measured TG and DSC data. Between 140 and 180 ◦C the apparent re- action rate increases nearly linearly. According to the Brack mechanism 13 this is mainly due to the reaction urea(l) HNCO(l) + NH (g). At this temperature, the vapor pressure of

3 Chemistry isocyanic acid is already more than 26 bar according to Equa- tion 20. Unless HNCO(l) reacts very fast to form biuret or triuret, it would evaporate, leaving no caloric contribution of HNCO(l) in a DSC experiment. A direct decomposition would lead to a net reaction

∆ ◦ urea(l) HNCO(g) + NH3(g) RH(160 C) = 153.8kJ/mol . Fig. 12 Proposed reaction scheme of urea decomposition (31) Physical According to Figure 11, we have a mass loss of 2.25 mg As a result of the aforementioned investigations, the authors (3.75·10−5 mol urea) for the temperature interval 140–180 ◦C. came up with the following reaction scheme (see Figure 12): The The DSC signal increases nearly linearly from 0.13 to 0.375 W/g urea decomposition is mainly driven by a thermodynamic equilib- within 20 min. This accounts for a total heat consumption of rium in the system consisting of urea, biuret, triuret and isocyanic

8 | 1–13 Page 9 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A

acid. Ammonia is produced by a kinetically limited process with the net reaction 2 urea(l) biu(l) + NH3(g). In analogy, a reaction to form triuret has been added: biu(l) + urea(l) triu(s) + NH3(g). Triuret then reacts further to form the solid deposits cyanuric acid and ammelide. Cyanuric acid sublimates for temperatures above 300 ◦C. This process is described here by a non-reversible surface process. However, it would make sense to describe it as an equilibrium process if reliable thermodynamic data for solid cyanuric acid would be available. For sake of com- pleteness, 6 global reaction steps for the decomposition of am- melide were added (see supplemental material). The kinetic parameters were adjusted manually until the main features of DSC and TG experiments were satisfied. A following

automated parameter optimization did not significantly improve Manuscript the results. Therefore, we use the manually fitted parameters for this publication (Table 3) in order to avoid pretending higher precision of the kinetic parameters. The biggest uncertainty is still the thermodynamic data of biuret and triuret. Fig. 14 TG of cyanuric acid (cylinder crucible, 2 K/min) with simulated composition of the mixture 6 Simulation results The quality of the proposed reaction mechanism can be evalu- The experiments with cyanuric acid are well modeled by the ated by comparison of the experimental data with the simulation Accepted sublimation reaction. The DSC signal (Fig. 13) and TG data results. Here, we will give results for some of the experiments Creative Commons Attribution-NonCommercial 3.0 Unported Licence. ((Fig. 14) show good agreement. For the plate-type crucible, the listed in Table 1. The other cases will be shown in supplemental sublimation is shifted correctly by ca. 40 K to lower temperatures. material. In reality, the dependency on the diameter of the crucible is less One of the consequences of the equilibrium processes involv- pronounced than in the simulation, because the surface area of ing isocyanic acid is that the model becomes sensitive to the the solid sample does not decrease linearly with the size of the gas-phase concentration of HNCO(g). This has not been taken

crucible. However, if the model is applied to thin layers of de- Physics into account during reactor model development when a zero- posits of cyanuric acid, it should be valid that the rate of sublima- dimensional approach was chosen to describe TG experiments. tion is proportional to the size of the deposit. The concentration gradient in the gas phase should be modeled

This article is licensed under a as well. To guarantee a sufficient dilution of the evaporating HNCO(g), the gas-phase volume was assumed to be a cylinder Triuret with the diameter of the crucible and a height of 2 m. Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM.

Cyanuric acid Chemical Chemistry

Fig. 15 DSC of triuret (cylinder crucible) Physical

A comparison of experimental data and simulation results for triuret decomposition is shown in Figures 15 and 16. In both Fig. 13 DSC of cyanuric acid (cylinder crucible) diagrams, two main decomposition steps can be identified: a step involving biuret for temperature between 190 and 230 ◦C and

1–13 | 9 Physical Chemistry Chemical Physics Page 10 of 14 View Article Online DOI: 10.1039/C9CP01529A

Table 3 Kinetic parameters of the proposed reaction scheme of urea decomposition

Reaction Ak (in SI units) βk Ea,k (in kJ/mol) Notes Homogeneous phase: urea(l) + HNCO(l) biu(s) 1·10−4 0 0 These three reactions feature HNCO(l) urea(l) + HNCO(l) biu(l) 1·10−4 0 0 even though the net product of the biu(l) + HNCO(l) triu(s) 1·10−4 0 0 reverse reaction is HNCO(g). · 2 triu(s) cya(s) + NH3(g) 1.2 10 0 45 · 1 triu(s) ammd(s) + H2O(g) 3 10 0 45 Surface: · 0 2 urea(l) biu(l) + NH3(g) 3.5 10 0 99 · 2 biu(l) + urea(l) triu(s) + NH3(g) 2 10 0 116.5 cya(s) cya(g) 3·104 0 141.3 Phase change (surface): · −2 H2O(g) H2O(l) 8.6 10 0.5 0 The condensation-evaporation urea(g) urea(l) 4.7·10−2 0.5 0 equilibria are based on HNCO(g) HNCO(l) 5.5·10−2 0.5 0 Herz-Knudsen model (Eq. 11). − urea(l) urea(s) 1·10 6 0 0 Adjusted to DSC signal of urea melting. Manuscript biu(l) biu(s) 1·10−6 0 0

because thermodynamically controlled equilibrium processes are more dominant than kinetically controlled reactions.

Biuret Accepted Creative Commons Attribution-NonCommercial 3.0 Unported Licence. Physics This article is licensed under a

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. Fig. 16 TG of triuret (cylinder crucible, 2 K/min) with simulated composi- tion of the mixture Chemical

the sublimation of cyanuric acid. The simulation broadens the first step to lower temperatures. This is because the TG data up to 200 ◦C have been used to adjust thermodynamic data of triuret on Fig. 17 DSC of biuret (cylinder crucible) the assumption of a equilibrium in the urea-biuret-triuret-HNCO ensemble. However, this does not account for the mass loss due to Unfortunately, no good agreement could yet be achieved for the kinetically limited reactions that release ammonia. Therefore, biuret (Figures 17 and 18). The interplay of thermodynam-

the total mass loss using the full mechanism is overestimated and ics and reaction kinetics does not resolve the effects in Fig. 3. Chemistry shifted to lower temperatures. Here, a better agreement could There should be tree decomposition steps: the eutectic urea- only be achieved by improving the thermodynamic data of biuret biuret phase, the biuret-triuret phase and the sublimation of cya- and triuret. nuric acid. For the same reason as the triuret decomposition is The simulation of TG experiments (Fig. 16) also gives an in- overestimated at lower temperatures, the triuret production is sight to the composition of the mixture in the crucible. No signifi- now underestimated in the first reaction step. Thus, the triuret cant amounts of liquid urea seem to be formed. Therefore, biuret decomposition does not contribute much to the TG signal in the ◦

also stays solid until complete decomposition at around 230 C. simulation, even though triuret is present at the expected temper- Physical Since the first decomposition step is slightly overestimated, the ature range 210–230 ◦C. Thermodynamic data should shift here amount of cyanuric acid is then underestimated. Ammelide for- the equilibrium of urea-biuret-triuret-HNCO towards triuret. Bet- mation is matched well when assuming that production of cya- ter agreement was achieved for the 10 K/min experiment when nuric acid and ammelide are in the ratio 4:1. The TG results decomposition is shifted to higher temperatures and the two de- show even better agreement for a temperature ramp of 10 K/min composition steps also merge in reality.

10 | 1–13 Page 11 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A Manuscript

Fig. 18 TG of biuret (cylinder crucible, 2 K/min) with simulated composi- Fig. 20 DSC of urea (cylinder crucible) tion of the mixture Accepted Creative Commons Attribution-NonCommercial 3.0 Unported Licence. Physics This article is licensed under a Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM.

Fig. 21 TG of urea (cylinder crucible, 2 K/min) with simulated composi- Chemical Fig. 19 TG of biuret maintaining a constant temperature of 200 ◦C tion of the mixture

when all urea is consumed. The produced amounts of biuret and On the other hand, simulations of TG experiments with biuret triuret seem to be in the right order of magnitude, leading to cor- where heating was stopped at 195, 200 and 210 ◦C show very rect predictions for cyanuric acid and ammelide. The predicted good agreement (Fig. 19). It should be noted that these exper- results again become even better for the 10 K/min ramp, when iments used a 15 times higher initial mass of biuret. Thus, the the first and the second decomposition step merge also in experi- contribution of volumetric equilibrium reactions is increased over Chemistry ment. the ammonia producing surface reactions. 7 Conclusions Urea At a first glance, there still seems to be too much uncertainty. For practical simulations of SCR systems, we are mainly interested The previous mechanism by Brack et al. 13 was able to model the in deposits resulting from urea decomposition. Here, the situation three decomposition steps of urea at the right temperature inter- looks better than for biuret. Experiments show three decomposi- vals. However, the proposed reaction scheme is advantageous in tion steps: a step in liquid urea-biuret mixture, a biuret-triuret several ways: Physical step and the sublimation of cyanuric acid. The initiation of the First, it is stringently based on thermodynamics. Melting and decomposition is well matched in DSC and TG (Figures 20 and solidification of substances are inherent to the thermodynamic 21). The biuret-triuret step occurs in simulation again shifted to data. The phase change during biuret decomposition can be ex- lower temperatures, but there is a phase change at around 200 ◦C, plained by formation of an eutectic mixture. There is no need for

1–13 | 11 Physical Chemistry Chemical Physics Page 12 of 14 View Article Online DOI: 10.1039/C9CP01529A

an extra matrix species. Hm,i Molar enthalpy (J/mol) Second, it is consistent with DSC experiments. From these, we h Henry constant (-) can conclude that cyanuric acid does not decompose to isocyanic Kc,k Equilibrium constant in terms of concentrations acid directly, but it sublimates instead. Furthermore we can also (variable units) rule out direct decomposition of urea to gaseous ammonia and Kp,k Equilibrium constant (-) 3 isocyanic acid. KH Henry’s coefficient (mol/m /Pa) 2 Third, it scales correctly with crucible size and gradient of the kW Heat transfer coefficient (W/m /K) temperature ramp. Thus, we have to distinguish between pro- Mi (kg/mol) cesses that are controlled by thermodynamic equilibrium and ki- m Mass (kg) netically controlled surface reactions. ni Molar amount (mol) Last, but not least, the overall reaction scheme is much simpler. n˙ik Molar rate of species Si by reaction Rk (mol/s) In the tradition of Occam’s razor, it explains more effects with less Pj Phase assumptions compared to the Brack mechanism. PDSC DSC signal (W)

The reaction mechanism is not perfect yet. The main reason is p Pressure (Pa) Manuscript ⊖ the lack of data for the thermodynamic potentials of biuret and p Standard pressure (101325 Pa) triuret. Eventually, a liquid triuret species should be added, be- pi Partial pressure (Pa) vap cause it should be part of the eutectic mixture. Currently, triuret pi Vapour pressure (Pa) decomposition is shifted to lower temperatures. This causes a R Gas constant (8.31446 J/mol/K) coincidence of the first and second decomposition step for biuret. Rk Reaction 3 2 For urea decomposition the results are in good agreement with rk Molar rate of reaction (mol/m /s or mol/m /s) 14 the measurements by Wang et al. The dominant process for bi- Si Species ⊖ uret and triuret formation is the self-combination of urea, which S Standard entropy at 298.15 K (J/mol/K) Accepted leads to the desorption of ammonia. However, this is a bit unsatis- Sm,i Molar entropy (J/mol/K) Creative Commons Attribution-NonCommercial 3.0 Unported Licence. fying from a viewpoint of elementary reaction step mechanisms. T Temperature (K) ⊖ According to the DSC signal, this reaction step cannot involve T Standard temperature (298.15 K) HNCO(l). A more elegant explanation might be a route through t Time (s) + – 3 an ionic reaction urea NH4 + OCN as suggested by Kieke Vj Volume of phase Pj (m ) 40 3 et al. Vm,i Molar Volume (m /mol) Physics To summarize, the authors propose a new reaction mechanism Xi Mol fraction (-) for the urea decomposition, which is closer to the real processes Yi Mass fraction (-) than the previously published mechanisms. We want to empha- αc Accumulation factor for condensation (-) This article is licensed under a size the importance of thermodynamics in modeling of deposit βk Temperature exponent ∆ ⊖ formation in SCR systems. Rk G Reaction free enthalpy (J/mol) ⊖ ∆fH Standard enthalpy of formation at 298.15 K (J/mol)

Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. Conflicts of interest ⊖ ∆solH Standard enthalpy of solution at 298.15 K (J/mol) µ There are no conflicts to declare. i Chemical potential of species Chemical µ0 Chemical potential of undiluted species Acknowledgements i νik Stoichiometric coefficient The authors kindly acknowledge financial support by German ν˜ik Concentration exponent 3 Research Foundation (Deutsche Forschungsgemeinschaft, DFG) ρi Density (kg/m ) through project 237267381 – TRR 150. DFG is also ac- knowledged for financing thermogravimetrical analysis equip- ment within project INST 121384/70-1. Furthermore we thank References Steinbeis GmbH & Co. KG für Technologietransfer (STZ 240 Reaktive Strömung) for a cost-free license of DETCHEM™. 1 Regulation (EC) No 715/2007 of the European Parliament and Chemistry of the Council of 20 June 2007 on type approval of motor vehi- Nomenclature cles with respect to emissions from light passenger and commer- A Cross-sectional area of the crucible (m2) cial vehicles (Euro 5 and Euro 6) and on access to vehicle repair 3 2 and maintenance information (Text with EEA relevance), 2007, Ak Pre-exponential factor (mol/m /s or mol/m /s) http://data.europa.eu/eli/reg/2007/715/oj. ai Chemical activity coefficient (-) ani Coefficient of NASA polynomials 2 F. Birkhold, U. Meingast, P. Wassermann and 3 Physical ci Concentration (mol/m ) O. Deutschmann, Applied Catalysis B-Environmental, 2007, ⊖ 3 70, 119–127. ci Reference concentration at normal pressure (mol/m ) cp,i Molar heat capacity (J/mol/K) 3 M. Börnhorst and O. Deutschmann, International Journal of Ea,k Activation energy (J/mol) Heat and Fluid Flow, 2018, 69, 55–61. H Enthalpy (J) 4 M. Börnhorst, S. Langheck, H. Weickenmeier, C. Dem,

12 | 1–13 Page 13 of 14 Physical Chemistry Chemical Physics View Article Online DOI: 10.1039/C9CP01529A

R. Suntz and O. Deutschmann, Chemical Engineering Journal, Chemie, 1938, 236, 200–208. 2018. 29 C. L. Speyers, American Journal of Science, 1902, 14, 293– 5 L. Stradella and M. Argentero, Thermochimica Acta, 1993, 302. 219, 315–323. 30 L. A. Pinck and M. A. Kelly, Journal of the American Chemical 6 A. G. Koryakin, V. A. Gal’perin, A. N. Sarbaev and A. I. Society, 1925, 47, 2170–2172. Finkel’shtein, Zh. Org. Khim., 1971, 972–977. 31 F.-M. Lee and L. E. Lahti, Journal of Chemical & Engineering 7 H. L. Fang and H. F. M. DaCosta, Applied Catalysis B: Environ- Data, 1972, 17, 304–306. mental, 2003, 46, 17–34. 32 VDI-Wärmeatlas: mit 320 Tabellen, ed. VDI, Springer Vieweg, 8 P. M. Schaber, J. Colson, S. Higgins, D. Thielen, B. Anspach Berlin, Heidelberg, 11th edn, 2013. and J. Brauer, Thermochimica Acta, 2004, 424, 131–142. 33 Lange’s Handbook of chemistry, ed. J. A. Dean and N. A. Lange, 9 A. Lundström, B. Andersson and L. Olsson, Chemical Engineer- McGraw-Hill, New York, 13th edn, 1985. ing Journal, 2009, 150, 544–550. 34 Kaye and Laby, Tables of Physical & Chemical Constants (16th 10 M. Eichelbaum, R. J. Farrauto and M. J. Castaldi, Applied edition 1995). 3.2 Properties of inorganic compounds. Version Catalysis B: Environmental, 2010, 97, 90–97. 1.0 (2005), Kaye & Laby Online, 2005. Manuscript 11 A. M. Bernhard, D. Peitz, M. Elsener, A. Wokaun and 35 ICSC Database - International Labour Organisation, O. Kröcher, Applied Catalysis B: Environmental, 2012, 115- 2017, http://www.ilo.org/safework/info/ 116, 129–137. publications/WCMS_113134/lang--en/index.htm. 12 V. Ebrahimian, A. Nicolle and C. Habchi, AIChE Journal, 2012, 36 F. L. Smith and A. H. Harvey, Chemical Engineering Progress, 58, 1998–2009. 2007, 103, 33–39. 13 W. Brack, B. Heine, F. Birkhold, M. Kruse, G. Schoch, S. Tis- 37 N. Borduas, B. Place, G. R. Wentworth, J. P. D. Abbatt and

cher and O. Deutschmann, Chemical Engineering Science, J. G. Murphy, Atmos. Chem. Phys., 2016, 16, 703–714. Accepted 2014, 106, 1–8. 38 http://roempp.thieme.de/. Creative Commons Attribution-NonCommercial 3.0 Unported Licence. 14 D. Wang, N. Dong, S. Hui and Y. Niu, FUEL, 2019, 242, 62–67. 39 A. I. Voskov, T. S. Babkina, A. V. Kuznetsov and I. A. Uspen- 15 O. Deutschmann, S. Tischer, S. Kleditzsch, V. Janardhanan, skaya, Journal of Chemical and Engineering Data, 2012, 57, C. Correa, D. Chatterjee, N. Mladenov, H. D. Minh, H. Karad- 3225–3232. eniz, M. Hettel, V. Menon and A. Banerjee, DETCHEM Soft- 40 M. L. Kieke, J. W. Schoppelrei and T. B. Brill, Journal of Phys- ware package, 2.7 ed., 2018, www.detchem.com. ical Chemistry, 1996, 100, 7455–7462. Physics 16 A. Burcat, Burcat’s Thermodynamic Data, http: //garfield.chem.elte.hu/Burcat/burcat.html. 17 T. Huthwelker and T. Peter, Journal of Chemical Physics, 1996, This article is licensed under a 105, 1661–1667. 18 P. Deuflhard, E. Hairer and J. Zugck, Numerische Mathematik, 1987, 51, 501–516. Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. 19 A. Roine, Outokumpu HSC Chemistry for Windows Chemi-

cal Reaction and Equilibrium Software with Extensive Thermo- Chemical chemical Database, 2002. 20 G. Y. Kabo, E. A. Miroshnichenko, M. L. Frenkel, A. A. Kozyro, V. V. Simirskii, A. Krasulin, V. Vorob’eva and Y. Lebedev, Bull. Acad. Sci. USSR, Div. Chem. Sci., 1990, 662–667. 21 http://webbook.nist.gov/. 22 A. A. Kozyro, M. L. Frenkel, A. P. Krasulin, V. V. Simirskii and G. Y. Kabo, Zhurnal Fizicheskoi Khimii, 1988, 62, 1752–1756.

23 R. C. Mann and C. K. Frederig, J. Chem. Soc. (Resumed) Royal Chemistry Society of Chemistry, 1943, 603–606. 24 A. A. Kozyro, G. Y. Kabo, T. V. Soldatova, V. V. Simirskii, V. I. Gogolinskii, A. P. Krasulin and N. M. Dudarevich, Zhur. Fiz. Khim., 1992, 66, 2583–2590. 25 H. G. M. DeWit, C. G. DeKruif and J. C. Van Miltenburg, J. Chem. Thermodynam., 1983, 15, 891–902.

26 V. D. Selivanov, V. M. Karlik and V. I. Zagranicnyi, Russ. J. Physical Phys. Chem. (Engl. Transl.), 1973, 47, 476–477. 27 D. Gratzfeld and M. Olzmann, Chemical Physics Letters, 2017, 619, 219–224. 28 M. Linhard, Zeitschrift für anorganische und allgemeine

1–13 | 13 Physical Chemistry Chemical Physics Page 14 of 14

View Article Online urea(g) HNCO(g) DOI: 10.1039/C9CP01529Acya(g)

HNCO(l) urea(l) biu(l) Manuscript

urea(s) biu(s) triu(s) cya(s) Accepted A new reaction mechanism of urea decomposition is proposed after a strict treatment of thermodynamics in all phases. Creative Commons Attribution-NonCommercial 3.0 Unported Licence. Physics This article is licensed under a Chemical Open Access Article. Published on 08 July 2019. Downloaded 7/8/2019 1:28:28 PM. Chemistry Physical