Next Article in Journal
Methane Combustion Kinetics over Palladium-Based Catalysts: Review and Modelling Guidelines
Previous Article in Journal
A Comprehensive Review of Fine Chemical Production Using Metal-Modified and Acidic Microporous and Mesoporous Catalytic Materials
Previous Article in Special Issue
Cobalt Nanoparticles Supported on TiO2 for Highly Selective Formation of N-Benzylideneanilines from Nitroarenes and Benzaldehyde via Reductive Imination Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Synthesis of Calcium Lactate through the Reaction of Glycerol with Calcium Hydroxide Catalyzed by Bimetallic AuCu/SiO2 Nanocatalysts

Faculty of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(5), 318; https://doi.org/10.3390/catal14050318
Submission received: 18 March 2024 / Revised: 8 May 2024 / Accepted: 9 May 2024 / Published: 11 May 2024

Abstract

:
Bimetallic AuCu/SiO2 nanosized catalysts were prepared using the wet chemical reduction technique. From among Au0.1–1.5Cu10/SiO2 catalysts, the Au0.5Cu10/SiO2 catalyst gave the highest yield of calcium lactate of 87% at a glycerol conversion of 96% when the reaction of glycerol with calcium hydroxide at a mole ratio of calcium hydroxide to glycerol of 0.8:1 was conducted under an anaerobic atmosphere at 200 °C for 2 h. The interactions between metallic Au0 and Cu0 nanoparticles facilitate calcium lactate formation. The simulation of glycerol consumption rate with an empirical power-function reaction kinetics equation yielded a reaction activation energy of 44.3 kJ∙mol−1, revealing that the catalytic reaction of glycerol with calcium hydroxide to calcium lactate can be conducted by overcoming a mild energy barrier. The synthesis of calcium lactate through the catalytic reaction of glycerol with calcium hydroxide on a bimetallic AuCu/SiO2 nanosized catalyst under a safe anaerobic atmosphere is an alternative to the conventional calcium lactate production technique through the reaction of expensive lactic acid with calcium hydroxide.

Graphical Abstract

1. Introduction

With the depletion of fossil fuels and the deterioration of the environment by CO2 emissions, the utilization of green and sustainable biomass and its related high value-added fine chemicals has gained much more attention worldwide. Glycerol, a byproduct in biodiesel production, has been produced at a large scale; around 1/10 of total biodiesel production capacity [1]. In addition to this, glycerol is also available in surfactant production from vegetable oils and animal fats. The worldwide glycerol market was estimated at about USD 615 million in 2022 and is expected to reach USD 954 million by 2029 [2]. The high glycerol production capacity makes glycerol conversion to high-value-added chemicals practical and beneficial.
Glycerol molecules of a polyhydroxy nature are multifunctional and can undergo catalytic reactions, such as hydrogenolysis, oxidation, and dehydration, to produce propanediol [3], glyceric acid, glycolic acid [4], dihydroxyacetone [5], glycerol carbonate [6], lactic acid [7,8,9], and calcium lactate [10]. From among them, calcium lactate can be used as a calcium supplement, antimicrobial agent, and in a variety of food additives in food, medicine, and feed fields.
Calcium lactate is commercially produced through the neutralization reaction of lactic acid with calcium hydroxide. For the raw material lactic acid, numerous studies have disclosed that noble metal catalysts, such as Au [11,12], Pt [13], and AuPt [14], can catalyze glycerol oxidation with gaseous O2 to lactic acid at reaction temperatures in a range of 60–100 °C, yielding lactic acid selectivities of 16%, 92.3%, 69.3%, and 83% at glycerol conversions of 15%, 82.4%, 100%, and 100%. Although noble metal catalysts can catalyze glycerol oxidation to lactic acid at mild reaction temperatures of around 95 °C, lower lactic acid selectivity at higher glycerol conversion rates and a higher noble metal catalyst cost limit their practical applications in lactic acid production.
Researchers have recently reported that non-precious metals exhibit impressively catalytic activities in hydrothermal glycerol conversion to lactic acid in alkaline aqueous solutions under an anaerobic atmosphere. It was reported that the glycerol conversion and lactic acid selectivity on the Cu/ZIF-8 catalyst were 97.5% and 84.2% when the glycerol conversion was conducted in sodium hydroxide aqueous solutions under an anaerobic N2 atmosphere at 230 °C for 8 h [15]. When the synthesis of lactic acid through hydrothermal glycerol conversion in sodium hydroxide aqueous solutions was conducted on a 20CuHT catalyst bed at 240 °C, the glycerol conversion and lactic acid yield were around 96.5% and 64%, respectively [16]. It was also reported that when the hydrothermal conversion of glycerol to lactate was conducted in calcium hydroxide aqueous solutions at 230 °C, the lactate selectivity was above 91% at a glycerol conversion of around 93% [10].
Cobalt-derived catalysts were also utilized in the conversion of glycerol to lactic acid in sodium hydroxide aqueous solutions at a higher reaction temperature of 250 °C [17]. At the comparable glycerol conversion of around 53.5%, the selectivities of lactic acid reached 90% for Co3O4/CeO2, 78% for Co3O4/ZrO2, and 68% for Co3O4/TiO2. Furthermore, Ye et al. [18,19] found that CaO and CaO/CuO catalysts could effectively catalyze the conversion of glycerol to lactic acid with a yield of 40.8% at a glycerol conversion of 97.8% and after esterification with methanol, the methyl lactate yield reached 57% at a glycerol conversion of 94%.
Active components and surface acidic/basic performance play important roles in catalytic glycerol conversion to lactic acid [10,15,16,17]. Considering that noble metal catalysts can catalyze glycerol conversion at a mild reaction temperature, whereas copper-based catalysts exhibit higher lactic acid selectivity at a higher temperature, we suggest that combining non-precious Cu with noble Au could produce good catalytic performance in glycerol conversion to lactate in an alkaline aqueous solution under anaerobic atmosphere.
In our present work, SiO2-nanoparticle-supported bimetallic AuCu nanocatalysts were prepared using a wet chemical reduction technique. The catalytic activity of the bimetallic AuCu/SiO2 nanocatalyst in the synthesis of calcium lactate through the reaction of glycerol with calcium hydroxide under an anaerobic atmosphere was assessed. And a power-function type reaction kinetics equation for glycerol consumption rate was simulated to assess the impacts of reaction parameters. Our present work reveals that calcium lactate can be effectively synthesized through the direct reaction between biomass glycerol and Ca(OH)2 catalyzed by the bimetallic AuCu/SiO2 catalyst, which is a sustainable methodology for the synthesis of calcium lactate instead of the conventional production technique through the reaction of expensive lactic acid with calcium hydroxide.

2. Results and Discussion

2.1. XRD Analysis

The X-ray diffractograms of Au1/SiO2, Cu10/SiO2, and Au0.1–1.5Cu10/SiO2 catalysts are illustrated in Figure 1. For all catalysts, wide XRD peaks centering at 2θ = 22.3° are ascribed to that of amorphous SiO2 nanoparticle support.
The XRD peaks of the Cu10/SiO2 catalyst centered at 2θ = 43.3, 50.5, and 74.3° are assigned to the (1 1 1), (2 0 0), and (2 2 0) crystal planes of metallic Cu0 (JCPDS 04-0836), respectively. For Au0.1–1.5Cu10/SiO2 catalysts, their XRD peaks at 2θ = 43.3, 50.5, and 74.2° are also assigned to the (1 1 1), (2 0 0), and (2 2 0) crystal planes of metallic Cu0 (JCPDS 04-0836). Calculated with the utilization of the Scherrer equation, crystallite sizes of metallic Cu (1 1 1) in the Au0.1–1.5Cu10/SiO2 catalysts decreased from 18.4 to 12.1 nm with the increase in Au content from 0% to 1.5 wt%. The presence of the Au component inhibited the crystal growth of the Cu0 component.
The XRD peaks of the Au1.0/SiO2 catalyst centered at 2θ = 38.1, 44.4, 64.6, and 77.5° are ascribed to (1 1 1), (2 0 0), (2 2 0), and (3 1 1) crystal planes of face-centered cubic Au0 (JCPDS 99-0056), respectively. For the Au0.1–1.5Cu10/SiO2 catalysts, at lower Au contents of less than 0.8 wt%, there were no XRD peaks for metallic Au0 detected, whereas with larger Au contents of 0.8 wt%, 1.0 wt%, and 1.5 wt%, only XRD peaks ascribed to the (1 1 1) crystal plane of metallic Au0 were detected, and XRD peaks centered at 38.3, 38.2, and 38.2°, respectively. As compared to the Au1.0/SiO2 catalyst, the positive shift in XRD peaks (1 1 1) of AuCu/SiO2 catalysts should be caused by interactions between Cu and Au components.
According to the Scherrer equation, the crystallite sizes of Au (1 1 1) of Au0.8Cu10/SiO2, Au1.0Cu10/SiO2, Au1.5Cu10/SiO2, and Au1.0/SiO2 catalysts were 6.7, 9.9, 12.2, and 18.1 nm, respectively. The results reveal that Au components in catalysts are at a nanosized scale. For the same Au content, the crystallite size of the Au component in the Au1.0Cu10/SiO2 catalyst is obviously less than that in the Au1.0/SiO2 catalyst, which could be interpreted as the interactions between Cu and Au components inhibiting the crystal growth of the metallic Au0 nanoparticles.

2.2. TEM and HRTEM Analyses

TEM images show that particle sizes of SiO2 support are in a range of 9–24 nm and center at 13 nm (Figure 2a–d). Average particle sizes and particle size distributions of metallic Cu0 and Au0 nanoparticles of the Cu10/SiO2 and Au1/SiO2 catalysts are 8.2, 5.3–12.3 and 3.4, 1.6–6.6 nm, respectively (Figure 2a,b). Average particle sizes and particle size distributions of metallic Cu0 and Au0 nanoparticles of the Au0.1Cu10/SiO2 and Au0.5Cu10/SiO2 are 8.2, 5.4–11.4; 2.9, 1.4–4.6 and 7.5, 5.1–10.7; and 2.8, 1.1–4.7 nm (Figure 2c,d). TEM and HRTEM images clearly disclose that Cu and Au components are present at a nanosized scale in catalysts.
The lattice fringes of Au1/SiO2 and Au0.1Cu10/SiO2 catalysts show that gold and copper components are present in metallic states (Figure 2e,f). The small-sized metallic Cu and Au nanoparticles were anchored and dispersed on the surfaces of SiO2 nanoparticles. The adjacent Cu and Au nanoparticles may have interactions.

2.3. XPS Analysis

Figure 3a shows that the binding energies of Si2p of SiO2 support, Au1.0/SiO2, Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts decrease from 103.7 to 103.2 eV with the introduction of Cu and Au components. The obvious shift of Si2p with the introduction of Cu and Au components reveals that there are interactions between SiO2 support and metallic Cu0 and Au0 components. Considering that the electronegativity of Si4+ (>14.15) is much larger than that of Au atoms (2.54) and Cu atoms (1.9) [20,21], the negative shift in Si2p indicates that electrons of Au and Cu components probably transfer to silicon (Si4+) components.
For the Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts, there were no characteristic satellite Cu2p peaks around 944 eV for Cu2+ observed (Figure 3b), which indicates that copper components in the above-mentioned catalysts are present at metallic Cu0 and/or Cu+ states [22]. Binding energies of Cu2p1/2 and Cu2p3/2 of the Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts were 952.7, 932.6; 952.4, 932.5; 952.2, and 932.2 eV. The addition of the Au component makes the Cu2p peaks shift negatively.
Binding energies of Au4f5/2 and Au4f7/2 of the Au1.0/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts were 87.2, 83.6; 88.2, 84.1; and 87.6, 83.8 eV (Figure 3c). The presence of metallic Cu0 in catalysts makes the Au4f shift positively. To reveal surface chemical valences of Au0 nanoparticles in Au1.0/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts, their XPS peaks were deconvoluted by utilizing the Gaussian–Lorentzian bands with peak positions at ca.87.2 and 83.6 eV for Au4f5/2 and Au4f7/2 of Au0 and ca. 88.2 and 84.2 eV for Au4f5/2 and Au4f7/2 of Auδ+ [23,24]. The deconvoluting results show that the area ratios of Auδ+/Au0 for Au4f5/2 and Au4f7/2 of Au1.0/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts were 0.03:1, 0.05:1; 9.41:1, 4.61:1; and 0.62:1, 0.55:1, respectively. The shifts in Au and Cu peaks reveal that there are interactions between Au0 and Cu0 nanoparticles in AuCu/SiO2 catalysts through the electron transfer from the Au component to the Cu component, which is consistent with that observed by XRD analysis. Interestingly, it was found that the shift extent of the Au4f of Au0.1Cu10/SiO2 was larger than that of Au0.5Cu10/SiO2. It could be explained because with a relatively lower Au content, Au0 nanoparticles could efficiently interact with Cu0 nanoparticles, whereas with a higher Au content, some of the Au0 nanoparticles are probably apart from the surfaces of the Cu0 nanoparticles.
To reveal the surface chemical valences of Cu0 nanoparticles in Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts, their XAES peaks were deconvoluted by utilizing the Gaussian–Lorentzian bands with peak positions at 916 eV for Cu+ and 918.7 eV for metallic Cu0, and 914.3 eV for other orbital electrons [25]. The deconvoluting results show that the area ratios of Cu0/Cu+ of Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts were 0.16:1, 0.26:1, and 0.67:1, respectively (Figure 3d). The electronegativities of Cu2+, Cu+, and Au0 were 10.28, 5.04, and 2.54, respectively [20,21]. The results disclose that the Au0 nanoparticles present in AuCu/SiO2 catalysts could transfer electrons to Cu2+ and even Cu+ components, which prevents the oxidation of Cu0 nanoparticles to some extent.
On the other hand, the atomic ratios of Au/Cu in Au0.1Cu10/SiO2 and Au0.5Cu10/SiO2 catalysts were 0.32:100 and 1.61:100, respectively. The XPS analysis showed that the atomic ratios of Au/Cu for Au0.1Cu10/SiO2 and Au0.5Cu10/SiO2 catalysts were 0.36:100 and 8.64:100. These results disclose that the surface compositions of the Au component are larger than the bulk ones. It is reasonable to suggest that metallic Au0 nanoparticles could anchor at surfaces of metallic Cu0 nanoparticles.

2.4. N2 Physisorption

As illustrated in Figure 4, nitrogen adsorption/desorption curves of representative samples show that the N2 adsorption/desorption isotherms belong to the III type. The SiO2 nanoparticle support has a large surface area of up to 221.2 m2∙g−1 (Table 1). For the Cu3–10/SiO2 catalysts, their surface areas decreased from 151.3 to 132.4 m2∙g−1 with the increase in Cu content from 3 wt% to 10 wt%. Their pore volumes and pore sizes were larger than those of the SiO2 support, respectively.
When the second metal, Au, was introduced, the surface areas of the Au0.1–1.5Cu10/SiO2 catalysts increased to 212.2 m2∙g−1 with the increase in Au content to 0.3 wt% and then decreased to 104.6 m2∙g−1 with the further increase in the Au content to 1.5 wt%. Both pore volume and pore size decreased with the increase in Au content in Au0.1–1.5Cu10/SiO2 catalysts. The small-sized Au nanoparticles easily choked catalyst pores, which lead to a decrease in the specific surface area of the Au1/SiO2 catalyst. When Cu nanoparticles were present in bimetallic CuAu/SiO2 catalysts, large-sized Cu nanoparticles reduced the choking to some extent, causing an increase in the surface area. However, the choking effect by Au nanoparticles increased with a higher content of Au nanoparticles, which decreased the specific surface area of the catalyst.
The Cu3–10/SiO2, Au1/SiO2, and Au0.1–1.5Cu10/SiO2 catalysts have large surface areas and pore sizes, which could supply a large amount of catalytic active sites and diminish mass transfer resistance in pores.

2.5. Catalytically Hydrothermal Conversion of Glycerol with Calcium Hydroxide to Calcium Lactate

2.5.1. Impact of Catalyst Composition

When the reaction of glycerol with calcium hydroxide at an initial mole ratio of calcium hydroxide to glycerol of 0.8:1 was catalyzed on Cu3–10/SiO2, Au1/SiO2, and Au0.1–1.5Cu10/SiO2 catalysts in aqueous solutions at 200 °C for 2 h, for the Cu3–10/SiO2 catalysts, the glycerol conversion, calcium lactate yield, and carbon balance increased with the increase in Cu content. From among Cu3–10/SiO2 catalysts, the Cu10/SiO2 catalyst gave the maximum glycerol conversion of 67% and a calcium lactate yield of 58% (Table 2). Larger metallic Cu content facilitates calcium lactate formation and inhibits byproduct formation.
The monometallic Au1/SiO2 catalyst exhibits poor catalytic performance in the reaction of glycerol with calcium hydroxide to calcium lactate, which leads to a much lower carbon balance value of 6%. This phenomenon is worthy of further investigation when metallic Au is solely used as the active component.
When metallic Au0 was introduced into the Cu10/SiO2 catalyst, glycerol conversions increased with the increase in Au content in bimetallic Au0.1–1.5Cu10/SiO2 catalysts. When Au contents were 1 wt% and 1.5 wt%, the glycerol conversions were above 99%. At a Au content of 0.5–1.0 wt%, the calcium lactate yields were above 86% with glycerol conversions ranging from 96% to 99%. As compared to the catalytic activity of the Cu3–10/SiO2 catalysts, the bimetallic AuCu/SiO2 catalyst exhibited high catalytic performance for calcium lactate synthesis through the hydrothermal reaction, starting from glycerol and calcium hydroxide at the reaction temperature of 200 °C.
It has been reported that small-sized CuO particles facilitate the conversion of glycerol to lactic acid [26]. However, when Cu-ZnO@C catalysts with metallic Cu particle sizes of 36.9, 67.8, and 93.6 nm catalyzed glycerol conversion to lactic acid, the Cu-ZnO@C catalyst with a metallic Cu particle size of 67.8 nm gave the highest lactic acid yield [27]. As compared to the Cu(16)/ZrO2 catalyst with a metallic Cu crystallite size of 44.5 nm, the Cu(16)HAP catalyst with a metallic Cu crystallite size of 33.8 nm exhibited a high glycerol conversion of 99% and a high lactic acid selectivity of 87% after reacting at 230 °C for 8 h [28]. Liu et al. [29] reported that the metallic Cu nanoparticle with a mean particle size of 11 nm exhibited the highest catalytic activity when Cu/SiO2 catalysts with mean metallic Cu particle sizes of 4–38 nm catalyzed xylitol hydrogenolysis. Generally, both the metallic Cu particle size and support type co-affect catalytic reactions.
In our present research, although the Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts had comparable metallic Cu nanoparticle sizes (around 8.0 nm), glycerol conversions on the Au0.1Cu10/SiO2 and Au0.5Cu10/SiO2 catalysts were 1.21 and 1.43 times those on the Cu10/SiO2 catalyst. The result clearly indicates that the interaction between Au0 and Cu0 nanoparticles improves the catalytic activity.

2.5.2. Impact of Reaction Temperature

Catalytic activities of the Au0.1–1.5Cu10/SiO2 catalysts in the reaction of glycerol with calcium hydroxide at an initial mole ratio of calcium hydroxide to glycerol of 0.8:1 to calcium lactate were investigated at 180, 200, and 230 °C for 2 h (Figure 5). When the Au content in AuCu/SiO2 catalysts was above 0.3 wt%, the glycerol conversion was more than 95% at the temperature of 200 °C. As the temperature was raised to 230 °C, on all the Au0.1–1.5Cu10/SiO2 catalysts, glycerol was almost completely converted (Figure 5a).
When bimetallic AuCu/SiO2 catalysts with Au content of 0.1–0.5 wt% catalyzed the reaction at 180 and 200 °C, the calcium lactate selectivity was above 90% (Figure 5b). When bimetallic AuCu/SiO2 catalysts with Au content of 0.8–1.5 wt% catalyzed the reaction at 180–230 °C, the calcium lactate selectivity was lower than 90%. The Au0.5Cu10/SiO2 catalyst gave the highest calcium yield at the reaction temperature of 200 °C.
At the reaction temperatures of 180–230 °C on the Au0.1–1.5Cu10/SiO2 catalysts, the selectivities of formate, acetate, and oxalate were below 2%, 6%, and 1%, respectively (Figure 5c–e). The AumCu10/SiO2 catalysts with a larger Au content gave slightly higher byproduct selectivity than those with a smaller Au content.
The catalysis results show that at a Au content of 0.5 wt%, the Au0.5Cu10/SiO2 catalyst is more profitable for calcium lactate formation through the catalytic reaction of glycerol with calcium hydroxide. As certified by the XPS analysis, at a Au content of 0.5 wt% or less, Au nanoparticles could efficiently interact with Cu nanoparticles, which improves the reaction of glycerol with calcium hydroxide to calcium lactate. At a larger Au content, isolated Au nanoparticles favorably lead to byproduct formation.

2.5.3. Impact of Calcium Hydroxide Content

When the reaction of glycerol with calcium hydroxide at various mole ratios was catalyzed by the Au0.5Cu10/SiO2 catalyst at 200 °C for 0.5–4 h, the glycerol conversion remarkably increased, whereas the calcium lactate selectivity slightly decreased with the increase in the initial mole ratio of calcium hydroxide to glycerol and the extension of the reaction time (Figure 6a,b). When the initial mole ratios of calcium hydroxide to glycerol were 0.6:1 and 0.8:1 at the reaction time of 2 h, glycerol conversions were 94% and 96%, respectively (Figure 6a). Selectivities of calcium lactate arrived at 91% (Figure 6b). Selectivities of calcium formate, acetate, and oxalate slightly increased with the increase in the initial mole ratio of calcium hydroxide to glycerol, but slightly decreased with the extension of the reaction time (Figure 6c–e).
Reaction routes in the catalytic reaction of glycerol with calcium hydroxide are illustrated in Scheme 1 according to the experimental results. The bimetallic Au0.5Cu10/SiO2 catalyst catalyzed the dehydrogenation of glycerol to glyceraldehyde [10,30,31], which is a slower reaction process because glyceraldehyde could not be identified under our experimental conditions. Glyceraldehyde could be catalytically dehydrated to form pyruvaldehyde by OH anions and the resultant pyruvaldehyde can be converted to calcium lactate [10,31,32]. The intermediates, glyceraldehyde and pyruvaldehyde, could also not be identified in our experiments, which indicates that the dehydration of glyceraldehyde and conversion of pyruvaldehyde are rapid reactions. The formation of oxalate, formate, and acetate could be through the decomposition of the intermediate and lactate catalyzed by OH anions [10]. The decrease in contents of oxalate, formate, and acetate with an extended reaction time is probably caused by their decomposition in an alkaline atmosphere.

2.5.4. Impacts of Glycerol Concentration and Catalyst Content

At initial glycerol concentrations of 0.5–2.0 mol∙L−1 and a mole ratio of calcium hydroxide to glycerol of 0.6:1, after reacting at 200 °C for 2 h on the Au0.5Cu/SiO2 catalyst, the glycerol conversions were above 94% at a calcium lactate selectivity of larger than 91% (Figure 7). When the glycerol concentration was further increased to 3.0 mol∙L−1, the glycerol conversion and calcium lactate selectivity remarkably declined to 70% and 63%, respectively. We suggest that the optimal concentration of glycerol is 2 mol∙L−1 for obtaining a higher calcium lactate yield under our present experimental conditions.
When the reaction of glycerol (2 mol∙L−1) with calcium hydroxide at an initial mole ratio of calcium hydroxide to glycerol of 0.6:1 was catalyzed by the Au0.5Cu10/SiO2 catalyst with catalyst contents of 1.84–9.2 g∙L−1 at 200 °C for 2 h, glycerol conversion and calcium lactate selectivity were obviously improved with the increase in catalyst content (Figure 8). With a catalyst content of 9.2 g∙L−1, the glycerol conversion reached 94% at a calcium lactate selectivity of 92%. Obviously, a high catalyst content is profitable for the catalytic reaction of glycerol with calcium hydroxide to calcium lactate.

2.6. Reaction Kinetics Simulation

To ascertain the impacts of the temperature, glycerol concentration, calcium hydroxide concentration, and catalyst content on the glycerol conversion rate, the reaction rate of glycerol consumption was simulated by utilizing an empirical power-function type reaction kinetics equation.
r A = d n A V d t = k C A a C B b C c a t c
k = A e x p ( E a / ( R T ) )
where k is the rate constant. a, b, and c are the reaction orders for glycerol concentration (CA, mol∙L−1), calcium hydroxide concentration (CB, g∙L−1), and catalyst concentration (Ccat, g∙L−1), respectively.
Although it is difficult to dissolve Ca(OH)2 in water, the produced lactate anion could rapidly react with Ca2+ to form easily soluble calcium lactate. The concentration of Ca(OH)2 is the whole mass concentration because the amount of Ca(OH)2 significantly affects the reaction. The catalyst concentration is the whole mass concentration.
Combing Equations (1) and (2) and taking the natural logarithm on both sides of the resultant equation yields Equation (3) in linear form.
ln r A = l n A E a / R T + a l n C A + b l n C B + c l n C c a t
Initial conversion rates of glycerol for the first 0.5 h under various reaction conditions were used for the reaction kinetics simulation by the linear regression method (Table S1). The simulated reaction kinetics equation with a correlation coefficient of 0.97 is listed as Equation (4). The higher correlation coefficient indicates that the reaction kinetic equation is well simulated using the reaction data under various reaction conditions.
r A = d n A / V d t = 247.87 exp ( 44,305.3 R T ) C A 0.23 C B 1.04 C c a t 0.7
The simulated reaction activation energy (Ea) is 44.3 kJ∙mol−1, which implies that glycerol conversion can be conducted by overcoming a mild energy barrier. According to the reaction orders, the impacts of concentrations of glycerol, calcium hydroxide, and catalyst on glycerol conversion rate are in the order of Ca(OH)2 > catalyst > glycerol.
Researchers reported that when the oxidation of glycerol to lactic acid was catalyzed by a CuO/C catalyst in a NaOH aqueous solution, the activation energy was 134.39 kJ·mol−1 [26]. The reaction activation energies for glycerol conversion to lactic acid on bimetallic CuAu1, CuAu2, CuAu3, and CuAu4 catalysts were 64.0, 53.4, 46.8, and 36.9 kJ∙mol−1, respectively [33]. Furthermore, Hausoul et al. [34] reported that in the hydrogenolysis of glycerol on the Pt/C catalyst, alkaline earth metal hydroxides improved the reaction. Therefore, we suggest that bimetallic CuAu and Ca(OH)2 facilitate glycerol conversion, which leads to a lower activation energy.

3. Experimental

3.1. Materials

Glycerol (99%), acetic acid (99.5%), oxalic acid (99.5%), lactic acid (85%), Cu(NO)3∙3H2O (99%), HAuCl4∙4H2O (Au, 47.8 wt%), hydrazine hydrate (85%), sodium hydroxide (NaOH, 96%), phosphoric acid (H3PO4, 85%), concentrated hydrochloric acid (HCl, 37%), disodium hydrogen phosphate (99%), n-butanol (99.5%), and methanol (HPLC, 99.8%) were supplied by the Dewei Chem. Reagent Co., Zhenjiang, China. Formic acid (88%) was supplied by the Nanjing Chem. Reagent Co., Nanjing, China. Silicon dioxide (average particle size, 13 nm) was supplied by the Hailong Nuclear Mater. Technol. Co., Zhenjiang, China.

3.2. Preparation of AuCu/SiO2 Nanocatalysts

Bimetallic AuCu/SiO2 nanocatalysts were prepared using the wet chemical reduction technique. First, 9.0 g of silicon dioxide was added to 40 mL of Cu(NO)3∙3H2O (0.4 mol∙L−1) aqueous solution. After the suspension was stirred at 40 °C for 1 h, a NaOH (4 wt%) aqueous solution was added dropwise to adjust the pH to 11.5. When the above-mentioned suspension was continuously stirred for 1 h, a prescribed volume of HAuCl4∙4H2O (1 wt%) aqueous solution was slowly dropped under stirring for 1 h and the pH value was maintained at 11.5. To reduce active components, hydrazine hydrate with a hydrazine hydrate/total metal ion molar ratio of 10:1 was added dropwise at 0.5 h. After reducing for 4 h, the resultant AuCu/SiO2 catalyst samples were separated by centrifugation at a rotation rate of 10,000 rpm for 15 min and washed with pure water to eliminate sodium salts. The residual concentrations of gold and copper components in the water phase were negligibly detected by ICP measurement, which indicates that gold and copper components could be completely loaded onto SiO2 support. The as-prepared AuCu/SiO2 nanocatalysts were kept in anhydrous ethanol to prevent oxidation by oxygen from the atmosphere. The catalysts were denoted as AumCun/SiO2 and m and n were the mass percentages of Au and Cu components in catalysts.

3.3. Characterization

Crystal structures and morphologies of the as-prepared AuCu/SiO2 nanocatalysts were examined on a powder X-ray diffractometer (D8 super speed Bruker-AEX Co. (Billerica, MA, USA), Cu Kα radiation at λ of 1.54056 Å) and a transmission electron microscope (JEM-2100, JEOL Ltd., Tokyo, Japan, 200 kV). Particle sizes of metallic Au0 and Cu0 nanoparticles on silicon dioxide nanoparticle supports were measured by combining both TEM and HRTEM analyses. The particle sizes were directly measured from TEM images of at least 50 individual particles.
X-ray photoelectron spectra of the AuCu/SiO2 nanocatalysts were acquired on an ESCALAB 250 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). Binding energies of Au, Cu, and Si components in catalysts were calibrated by referencing the C1s peak at 284.8 eV of contaminated carbon.
To determine surface areas and pore sizes of the AuCu/SiO2 nanocatalysts, their N2 adsorption–desorption isotherms were acquired by utilizing a NOVA 2000e physical adsorption apparatus at −196 °C (Quantachrome, Boynton Beach, FL, USA).

3.4. Catalysis

The synthesis of calcium lactate was catalyzed by the as-prepared AuCu/SiO2 nanocatalysts through a glycerol dehydrogenation reaction in a calcium hydroxide aqueous solution under a nitrogen atmosphere. The prescribed quantities of glycerol, calcium hydroxide, and catalysts were added to deionized water to obtain 100 mL of reaction mixture in a 1 L steel stainless autoclave. After a pure nitrogen (99.999%) stream was introduced into the reactor to replace air inside for at least 0.25 h, reaction temperatures were raised to prescribed values by heating the autoclave with the use of a 1200 W electric oven outside. As the reaction temperatures arrived at prescribed values, stirring started at 500 rpm, and the timing of the reaction was started.
After reacting for a prescribed time, the spent AuCu/SiO2 catalyst was filtrated from the reaction mixture at 80 °C. The filtrate was acidified with concentrated hydrochloric acid to a pH of around 1.0 and analyzed using the high-performance liquid chromatography(Agilent 1290 Infinity II, Agilent Technologies, Santa Clara, CA, USA), utilizing a reversed-phase C18 column (Wonda Cr act ODS-2, 5 μm, 4.6 × 250 nm) at 35 °C. H3PO4/Na2HPO4 (20 mmol∙L−1 Na2HPO4, acidified by H3PO4 to pH = 2.5) aqueous solution and methanol (v:v = 90:10) were mixed and utilized as the mobile phase with a volumetric flowing rate at 0.7 mL∙min−1. Products were identified by comparison with the retention times of standard samples, and their compositions were quantified by an external standard method. The quantity of unreacted glycerol was quantitatively analyzed by gas chromatography (Agilent Technologies 7890A, PEG-20M, Agilent Technologies, Santa Clara, CA, USA) with an FID detector using n-butanol as an internal standard. The calculation of product selectivity was based on the carbon number of the product molecules.

4. Conclusions

SiO2 nanoparticle-supported bimetallic AuCu (AuCu/SiO2) nanocatalysts were prepared using the wet chemical reduction technique. The metallic Au0 and Cu0 nanoparticles in AuCu/SiO2 catalysts experienced interactions by electron transfer from metallic Au to metallic Cu nanoparticles.
From among the Au0.1–1.5Cu10/SiO2 catalysts, the Au0.5Cu10/SiO2 and Au0.8Cu10/SiO2 catalysts gave high calcium lactate yields of 87% and 86% at glycerol conversion rates of 96% and 97% when the reaction of glycerol with calcium hydroxide at an initial mole ratio of calcium hydroxide to glycerol of 0.8:1 was conducted under anaerobic atmosphere at 200 °C for 2 h. The interaction between Cu and Au nanoparticles improved the synthesis of calcium lactate through the direct reaction between glycerol and Ca(OH)2 under an anaerobic condition.
The consumption rate of glycerol was well simulated with the empirical power-function type reaction kinetics, which gave the reaction activation energy (Ea) of 44.3 kJ∙mol−1 and the reaction orders of 0.23 for glycerol concentration, 1.04 for Ca(OH)2 concentration, and 0.7 for the Au0.5Cu10/SiO2 catalyst. The simulated reaction kinetics show that the synthesis of calcium lactate through the catalytic reaction of glycerol with calcium hydroxide could be efficiently carried out by overcoming a mild energy barrier.
The present research shows that the synthesis of calcium lactate can be efficiently and directly conducted through the catalytic reaction of glycerol with calcium hydroxide on a bimetallic AuCu/SiO2 nanosized catalyst under a safe anaerobic atmosphere. This new methodology could replace the conventional calcium lactate production technique through the neutralization reaction between expensive lactic acid and Ca(OH)2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14050318/s1, Table S1: Glycerol conversion rates under various reaction conditions catalyzed by the Au0.5Cu10/SiO2 catalyst.

Author Contributions

C.L.: Experimental research, Data analysis, Writing—original draft. X.C.: Experimental research, Data analysis. A.W.: Methodology, Conceptualization, Experimental research. H.Y.: Writing—review and editing, Funding acquisition, Supervision. Y.L.: Experimental research, Data analysis. Q.L.: Experimental research, Data analysis. J.G.: Experimental research, Data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the research funds from the Department of Liaoning Science and Technology, China (2021JH1/10400063, XLYC2204029).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Raman, A.A.A.; Tan, H.W.; Buthiyappan, A. Two-step purification of glycerol as a value added by product from the biodiesel production process. Front. Chem. 2019, 7, 774. [Google Scholar] [CrossRef] [PubMed]
  2. Global Info Research, Global Glycerol Market 2023 by Manufacturers, Regions, Type and Application, Forecast to 2029. Available online: https://www.globalinforesearch.com/search/glycerol (accessed on 26 January 2024).
  3. Agrawal, D.; Budakoti, M.; Kumar, V. Strategies and tools for the biotechnological valorization of glycerol to 1, 3-propanediol: Challenges, recent advancements and future outlook. Biotechnol. Adv. 2023, 66, 108177. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, S.; Xiao, Y.; Zhang, W.; Liao, S.; Yang, R.; Li, J.; Hu, C. Relay catalysis of copper-magnesium catalyst on efficient valorization of glycerol to glycolic acid. Chem. Eng. J. 2022, 428, 132555. [Google Scholar] [CrossRef]
  5. Liu, D.; Liu, J.-C.; Cai, W.; Ma, J.; Yang, H.B.; Xiao, H.; Li, J.; Xiong, Y.; Huang, Y.; Liu, B. Selective photoelectrochemical oxidation of glycerol to high value-added dihydroxyacetone. Nat. Commun. 2019, 10, 1779. [Google Scholar] [CrossRef] [PubMed]
  6. Keogh, J.; Deshmukh, G.; Manyar, H. Green synthesis of glycerol carbonate via transesterification of glycerol using mechanochemically prepared sodium aluminate catalysts. Fuel 2021, 310, 122484. [Google Scholar] [CrossRef]
  7. Pemmana, H.R.; Barnwal, P.K.; Uppaluri, R.V.S.; Peela, N.R. Selective aerobic-oxidation of glycerol to lactic acid over ruthenium-vanadium bimetallic catalysts. J. Ind. Eng. Chem. 2023, 124, 224–231. [Google Scholar] [CrossRef]
  8. Wang, C.; Zhang, X.; Li, J.; Qi, X.; Guo, Z.; Wei, H.; Chu, H. Gold nanoparticles on nanosheets derived from layered rare-earth hydroxides for catalytic glycerol-to-lactic acid conversion. ACS Appl. Mater. Interfaces 2021, 13, 522–530. [Google Scholar] [CrossRef]
  9. Razali, N.; Abdullah, A.Z. Production of lactic acid from glycerol via chemical conversion using solid catalyst: A review. Appl. Catal. A Gen. 2017, 543, 234–246. [Google Scholar] [CrossRef]
  10. Wang, A.; Xu, Q.; Yin, H.; Xiao, L.; He, Y. Synthesis of calcium lactate starting from glycerol and calcium hydroxide catalyzed by copper-based catalysts in aqueous solution. Mater. Sci. Eng. B 2023, 296, 116678. [Google Scholar] [CrossRef]
  11. Redina, E.A.; Kapustin, G.I.; Tkachenko, O.P.; Greish, A.A.; Kustov, L.M. Effect of ultra-low amount of gold in oxide supported bimetallic Au–Fe and Au–Cu catalysts on liquid-phase aerobic glycerol oxidation in water. Catal. Sci. Technol. 2021, 11, 5881–5897. [Google Scholar] [CrossRef]
  12. Sever, B.; Yildiz, M. Conversion of glycerol to lactic acid over Au/bentonite catalysts in alkaline solution. React. Kinet. Mech. Catal. 2020, 130, 863–874. [Google Scholar] [CrossRef]
  13. Zhang, C.; Wang, T.; Liu, X.; Ding, Y. Selective oxidation of glycerol to lactic acid over activated carbon supported Pt catalyst in alkaline solution. Chin. J. Catal. 2016, 37, 502–509. [Google Scholar] [CrossRef]
  14. Evans, C.D.; Douthwaite, M.; Carter, J.H.; Pattisson, S.; Kondrat, S.A.; Bethell, D.; Knight, D.W.; Taylor, S.H.; Hutchings, G.J. Enhancing the understanding of the glycerol to lactic acid reaction mechanism over AuPt/TiO2 under alkaline conditions. J. Chem. Phys. 2020, 152, 134705. [Google Scholar] [CrossRef] [PubMed]
  15. Qiu, K.; Shu, Y.; Zhang, J.; Gao, L.; Xiao, G. Effective and stable zeolite imidazole framework-supported copper nanoparticles (Cu/ZIF-8) for glycerol to lactic acid. Catal. Lett. 2022, 152, 172–186. [Google Scholar] [CrossRef]
  16. Moreira, A.B.F.; Nogueira, D.S.; Freitas, I.C.; Souza Júnior, R.L.; Chagas, C.A.; Souza, M.M.V.M.; Manfro, R.L. Hydrothermal transformation of glycerol to lactic acid in alkaline medium using Cu catalysts obtained from hydrotalcite-like precursors. Waste Biomass Valori. 2023, 14, 2335–2347. [Google Scholar] [CrossRef]
  17. Torres, S.; Palacio, R.; López, D. Support effect in Co3O4-based catalysts for selective partial oxidation of glycerol to lactic acid. Appl. Catal. A Gen. 2021, 621, 118199. [Google Scholar] [CrossRef]
  18. Chen, L.; Ren, S.; Ye, X.P. Lactic acid production from glycerol using CaO as solid base catalyst. Fuel Process. Technol. 2014, 120, 40–47. [Google Scholar] [CrossRef]
  19. Ren, S.; Ye, X.P.; Ayers, P.D. Tandem catalytic conversion of glycerol using solid catalysts followed by transesterification to produce alkyl lactate. RSC Adv. 2015, 5, 53230–53239. [Google Scholar] [CrossRef]
  20. Mann, J.B.; Meek, T.L.; Knight, E.T.; Capitani, J.F.; Allen, L.C. Configuration energies of the d-block elements. J. Am. Chem. Soc. 2000, 122, 5132–5137. [Google Scholar] [CrossRef]
  21. Feng, Y.; Zhang, F.; Xuan, L.; Zhu, Y. Ionic electronegativity. J. Dalian Inst. Light Ind. 1998, 17, 70–76. [Google Scholar]
  22. Yao, Y.; Wu, X.; Chen, B.; Tu, Z.; Gutiérrez, O.Y.; Cui, Y.; Wang, J.; Huang, J.; Xu, Y.; Sun, H.; et al. Copper-based catalyst confined in carbon nanocages reactor for condensed esters hydrogenation: Tuning of copper species by confined SiO2 and the methanol resistance. ACS Sustain. Chem. Eng. 2021, 9, 16270–16280. [Google Scholar] [CrossRef]
  23. Hernández-Ramírez, E.; Wang, J.A.; Chen, L.F.; Valenzuela, M.A.; Dalai, A.K. Partial oxidation of methanol catalyzed with Au/TiO2, Au/ZrO2 and Au/ZrO2-TiO2 catalysts. Appl. Surf. Sci. 2017, 399, 77–85. [Google Scholar] [CrossRef]
  24. Zhang, A.; Zhang, L.; Jing, G.; Zhang, H.; Wang, S.; Su, H.; Zeng, S. Promotion of Au3+ reduction on catalytic performance over the Au/CuOeCeO2 catalysts for preferential CO oxidation. Int. J. Hydrogen Energy 2018, 43, 10322–10333. [Google Scholar] [CrossRef]
  25. Wang, A.; Zhang, M.; Yin, H.; Liu, S.; Liu, M.; Hu, T. Direct reaction between silicon and methanol over Cu-based catalysts: Investigation of active species and regeneration of CuCl catalyst. RSC Adv. 2018, 8, 19317–19325. [Google Scholar] [CrossRef] [PubMed]
  26. Li, S.; Li, S.M.; Wang, Y.; Tang, C.; Qiu, L.; Yu, S. Selective oxidation of glycerol to lactic acid catalyzed by CuO/activated carbon and reaction kinetics. ACS Omega 2024, 9, 10583–10591. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, J.; Zhu, G.; Wang, X.; Mai, Y.; Chen, J. Selective catalytic conversion of glycerol to lactic acid over Cu-ZnO@C catalysts. Catal. Commun. 2023, 181, 106733. [Google Scholar] [CrossRef]
  28. Yin, H.X.; Zhang, C.; Yin, H.; Gao, D.; Shen, L.; Wang, A. Hydrothermal conversion of glycerol to lactic acid catalyzed by Cu/hydroxyapatite, Cu/MgO, and Cu/ZrO2 and reaction kinetics. Chem. Eng. J. 2016, 288, 332–343. [Google Scholar] [CrossRef]
  29. Liu, H.; Huang, Z.; Xia, C.; Jia, Y.; Chen, J.; Liu, H. Selective hydrogenolysis of xylitol to ethylene glycol and propylene glycol over silica dispersed copper catalysts prepared by a precipitation–gel method. ChemCatChem 2014, 6, 2918–2928. [Google Scholar] [CrossRef]
  30. Liu, L.; Ye, X.P. Simultaneous production of lactic acid and propylene glycol from glycerol using solid catalysts without external hydrogen. Fuel Process. Technol. 2015, 137, 55–65. [Google Scholar] [CrossRef]
  31. Wang, A.; Xu, Q.; Yin, H. Synthesis of lactic acid starting from glycerol catalyzed by CaO supported CuO and metallic Cu catalysts in Ca(OH)2 aqueous solution. React. Kinet. Mech. Catal. 2022, 135, 3205–3221. [Google Scholar] [CrossRef]
  32. Roy, D.; Subramaniam, B.; Chaudhari, R.V. Cu-based catalysts show low temperature activity for glycerol conversion to lactic acid. ACS Catal. 2011, 1, 548–551. [Google Scholar] [CrossRef]
  33. Shen, L.; Zhou, X.; Wang, A.; Yin, H.; Yin, H.X.; Cui, W. Hydrothermal conversion of high-concentrated glycerol to lactic acid catalyzed by bimetallic CuAux (x = 0.01–0.04) nanoparticles and their reaction kinetics. RSC Adv. 2017, 7, 30725–30739. [Google Scholar] [CrossRef]
  34. Beine, A.K.; Wang, X.; Vennewald, M.; Schmidt, R.U.S.; Glotzbach, C.; Palkovits, R.; Hausoul, P.J.C. On the effect of alkaline earth metal cations in the hydrogenolysis of glycerol over Pt/C—An experimental and theoretical study. ChemCatChem 2022, 14, e202101940. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Au1/SiO2, Cu10/SiO2, and bimetallic AuCu/SiO2 catalysts.
Figure 1. XRD patterns of the Au1/SiO2, Cu10/SiO2, and bimetallic AuCu/SiO2 catalysts.
Catalysts 14 00318 g001
Figure 2. TEM images of the (a) Cu10/SiO2, (b) Au1/SiO2, (c) Au0.1Cu10/SiO2, and (d) Au0.5Cu10/SiO2 catalysts and HRTEM images of (e) Au1/SiO2 and (f) Au0.1Cu10/SiO2 catalysts.
Figure 2. TEM images of the (a) Cu10/SiO2, (b) Au1/SiO2, (c) Au0.1Cu10/SiO2, and (d) Au0.5Cu10/SiO2 catalysts and HRTEM images of (e) Au1/SiO2 and (f) Au0.1Cu10/SiO2 catalysts.
Catalysts 14 00318 g002
Figure 3. XPS of (a) Si2p, (b) Cu2p, and (c) Au4f of the SiO2 support, Au1.0/SiO2, Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts and (d) Cu XAES of the Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts.
Figure 3. XPS of (a) Si2p, (b) Cu2p, and (c) Au4f of the SiO2 support, Au1.0/SiO2, Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts and (d) Cu XAES of the Cu10/SiO2, Au0.1Cu10/SiO2, and Au0.5Cu10/SiO2 catalysts.
Catalysts 14 00318 g003
Figure 4. Nitrogen adsorption/desorption curves of the SiO2 support and representative Cu/SiO2 and AuCu/SiO2 catalysts.
Figure 4. Nitrogen adsorption/desorption curves of the SiO2 support and representative Cu/SiO2 and AuCu/SiO2 catalysts.
Catalysts 14 00318 g004
Figure 5. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.1–1.5Cu10/SiO2 catalysts. Glycerol aqueous solution, 2 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.8:1; catalyst, 9.2 g∙L−1; reaction time, 2 h. (a) Glycerol conversion and (be) selectivities of calcium lactate, formate, acetate, and oxalate.
Figure 5. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.1–1.5Cu10/SiO2 catalysts. Glycerol aqueous solution, 2 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.8:1; catalyst, 9.2 g∙L−1; reaction time, 2 h. (a) Glycerol conversion and (be) selectivities of calcium lactate, formate, acetate, and oxalate.
Catalysts 14 00318 g005
Figure 6. Reaction of glycerol with calcium hydroxide at different ratios catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 2 mol∙L−1; reaction temperature, 200 °C; catalyst, 9.2 g∙L−1. (a) Glycerol conversion and (be) selectivities of calcium lactate, formate, acetate, and oxalate.
Figure 6. Reaction of glycerol with calcium hydroxide at different ratios catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 2 mol∙L−1; reaction temperature, 200 °C; catalyst, 9.2 g∙L−1. (a) Glycerol conversion and (be) selectivities of calcium lactate, formate, acetate, and oxalate.
Catalysts 14 00318 g006aCatalysts 14 00318 g006b
Scheme 1. Reaction routes in the catalytic reaction of glycerol with calcium hydroxide in an aqueous solution on bimetal AuCu/SiO2 catalysts [10,30,32].
Scheme 1. Reaction routes in the catalytic reaction of glycerol with calcium hydroxide in an aqueous solution on bimetal AuCu/SiO2 catalysts [10,30,32].
Catalysts 14 00318 sch001
Figure 7. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 0.5–3 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.6:1; catalyst, 9.2 g∙L−1; reaction time, 2 h at 200 °C. (a) Glycerol conversion and calcium lactate selectivity and (b) selectivities of calcium acetate, formate, and oxalate.
Figure 7. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 0.5–3 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.6:1; catalyst, 9.2 g∙L−1; reaction time, 2 h at 200 °C. (a) Glycerol conversion and calcium lactate selectivity and (b) selectivities of calcium acetate, formate, and oxalate.
Catalysts 14 00318 g007
Figure 8. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 2 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.6:1; catalyst, 1.84–9.2 g∙L−1; reaction time, 2 h at 200 °C. (a) Glycerol conversion and calcium lactate selectivity and (b) selectivities of calcium acetate, formate, and oxalate.
Figure 8. Reaction of glycerol with calcium hydroxide catalyzed by the Au0.5Cu10/SiO2 catalyst. Glycerol aqueous solution, 2 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.6:1; catalyst, 1.84–9.2 g∙L−1; reaction time, 2 h at 200 °C. (a) Glycerol conversion and calcium lactate selectivity and (b) selectivities of calcium acetate, formate, and oxalate.
Catalysts 14 00318 g008
Table 1. Structural properties of the SiO2, Cu/SiO2, Au/SiO2, and AuCu/SiO2 catalysts.
Table 1. Structural properties of the SiO2, Cu/SiO2, Au/SiO2, and AuCu/SiO2 catalysts.
CatalystsSBET (m2∙g−1)Vpore (cm3∙g−1)Dpore (nm)
SiO2221.40.5711.10
Cu3/SiO2151.31.1428.34
Cu5/SiO2141.01.0726.67
Cu10/SiO2132.40.7121.06
Au1/SiO2104.20.7528.16
Au0.1Cu10/SiO2151.90.7318.83
Au0.3Cu10/SiO2212.20.7315.05
Au0.5Cu10/SiO2177.80.6514.07
Au0.8Cu10/SiO2160.50.6012.76
Au1.0Cu10/SiO2105.00.6211.96
Au1.5Cu10/SiO2104.60.6011.91
Table 2. Catalytic performance of the SiO2, Cu/SiO2, Au/SiO2, and AuCu/SiO2 catalysts.
Table 2. Catalytic performance of the SiO2, Cu/SiO2, Au/SiO2, and AuCu/SiO2 catalysts.
CatalystsGlycerol Conversion
(%)
Product Selectivity
(%)
Yield
(%)
Carbon
Balance (%)
OAFALAAALA
Cu3/SiO2220.256081373.2
Cu5/SiO2440.117723480.1
Cu10/SiO2670.228715890.2
Au1/SiO2240.30.24115.5
Au0.1Cu10/SiO2810.229437699.2
Au0.3Cu10/SiO2900.229338498.2
Au0.5Cu10/SiO2960.229138796.2
Au0.8Cu10/SiO2970.228948695.2
Au1.0Cu10/SiO2990.218738691.2
Au1.5Cu10/SiO21000.228448490.2
Glycerol aqueous solution, 2.0 mol∙L−1; initial mole ratio of calcium hydroxide to glycerol at 0.8:1; catalyst, 9.2 g∙L−1; reaction time, 2.0 h; reaction temperature, 200 °C. OA, oxalate; FA, formate; LA, lactate; AA, acetate.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, C.; Cui, X.; Wang, A.; Yin, H.; Li, Y.; Lin, Q.; Guo, J. Direct Synthesis of Calcium Lactate through the Reaction of Glycerol with Calcium Hydroxide Catalyzed by Bimetallic AuCu/SiO2 Nanocatalysts. Catalysts 2024, 14, 318. https://doi.org/10.3390/catal14050318

AMA Style

Li C, Cui X, Wang A, Yin H, Li Y, Lin Q, Guo J. Direct Synthesis of Calcium Lactate through the Reaction of Glycerol with Calcium Hydroxide Catalyzed by Bimetallic AuCu/SiO2 Nanocatalysts. Catalysts. 2024; 14(5):318. https://doi.org/10.3390/catal14050318

Chicago/Turabian Style

Li, Changqing, Xinyue Cui, Aili Wang, Hengbo Yin, Yuting Li, Qiao Lin, and Junjie Guo. 2024. "Direct Synthesis of Calcium Lactate through the Reaction of Glycerol with Calcium Hydroxide Catalyzed by Bimetallic AuCu/SiO2 Nanocatalysts" Catalysts 14, no. 5: 318. https://doi.org/10.3390/catal14050318

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop