Next Article in Journal
Experimental Study on Repairing the Mechanical Characteristics of Oil-Contaminated Silty Clay in Ancient Dike with Modified Lime Mortar
Previous Article in Journal
Flexural Properties of Lattices Fabricated with Planar and Curved Layered Fused Filament Fabrication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Determination and Calculations of Mercury Vapor Concentration and Energy Released from Freshly Condensed Dental Amalgams Having Various Copper Percentages within the Alloy

Department of Mechanical Engineering, North Carolina A&T State University, Greensboro, NC 27411, USA
*
Author to whom correspondence should be addressed.
Materials 2023, 16(9), 3452; https://doi.org/10.3390/ma16093452
Submission received: 3 March 2023 / Revised: 21 April 2023 / Accepted: 25 April 2023 / Published: 28 April 2023
(This article belongs to the Section Metals and Alloys)

Abstract

:
Dental amalgam is an alloy consisting of a mixture of fine metallic powder of silver, tin, zinc, copper, and a trace amount of palladium in combination with about fifty percent elemental mercury that forms a matrix phase. Dental amalgams consisting of a high-copper content are the most common types of alloys currently utilized for the restoration of decayed, broken, and fractured posterior human teeth. The present research objective was primarily to improve the material properties by determining and analyzing the amount of mercury vapor released from dental amalgam received from eight different commercial brands. The mechanical hardness of the alloys was found to increase with an increase in copper content in the amalgam. The effect of copper addition on material aging was also studied. During the release of mercury vapor, the corresponding energies associated with the release of mercury vapor from each sample were determined for each successive measurement. The results indicated that increasing the copper content of the amalgam counters the release of mercury vapor from posterior teeth and improves the hardness properties.

1. Introduction

Dental amalgam material has been utilized by dentists for the restoration of mainly posterior teeth in humans since about 1833, when it was introduced in the United States from England [1]. There are other dental restorative materials than amalgams being used today, but dentists choose to use dental amalgam for posterior teeth restoration due to its higher relative strength and cost-effectiveness compared to composite resin [1,2]. In 2008, the use of dental amalgam was banned in countries such as Norway, Sweden, and Denmark due to concerns regarding mercury vapor toxicity [3]. In the United States, the American Dental Association (ADA) and the Government approved amalgam’s use in dentistry, as major manufacturers are motivated by sheer profit and dental amalgam’s affordability to customers [4]. Dental restorative material has been known to release mercury vapor at very high levels of concentration when triturated and during subsequent condensation as part of cavity preparation in human teeth [3]. The silver in the alloy is responsible for the setting expansion and creates good atomic bonding with tin, enabling an increase in strength and resistance to corrosion within the alloy [4]. More fundamentally, the silver renders the material adapted to high surface shine when polished after placement in the tooth-prepared cavity by the dentist. The tin usually causes setting contraction and gives the alloy its malleability property. Copper improves strength, minimizes corrosion and tarnishing, reduces creep, and improves marginal leakages. Zinc is the scavenger element that reduces oxidation and enables the removal of impurities in the alloy [5,6].
The classification of dental amalgam falls into two categories: (1) low-copper amalgam (about 6% Cu, conventional alloy), and (2) high-copper amalgam (>6% Cu) [6]. Low-copper amalgam has a composition of 55–60 wt. % silver, 20–29 wt. % tin, 6–8 wt. % copper, and <2 wt. % zinc [6]. The chemical formula of the amalgam is given as Ag3Sn (silver–tin γ) + Hg (elemental-mercury) → Ag3Sn (silver–tin γ phase) + Ag2Hg3 (silver–mercury γ1 phase) + Sn8Hg (tin–mercury γ2 phase) [6]. This takes place with a reduction in the size of the particle as the solubility of Sn and Ag in Hg is limited (0.35 and 0.6 wt. %, respectively) [4,7,8,9]. The silver precipitates out initially as Ag-Hg (γ1 phase), followed by tin in the form of the Sn-Hg (γ2 phase). The set amalgam consists of core γ particles surrounded by a matrix of γ1 and γ2 phases [10,11].
In low-copper alloys, the anodic potential of the Sn8Hg (γ2 phase) is greater than that of the Ag2Hg31 phase), and this leads to corrosion and resultant distribution of the γ2 phase; thus, forming the reaction products that develop a protective layer of oxide and hydroxide on the outer surface of the amalgam due to aging [12,13]. It has been established conclusively that this founding phenomenon helps in the deceleration of the rate of oxidation. Mercury in the Sn8Hg (γ2 phase) is released either as mercury vapor or as solutes in the saliva. Since the Sn8Hg (γ2 phase) occupies about 15% vol. of the mixed alloy as a continuous or near-continuous structural component, the corrosion loss of this component weakens the matrix and renders it more susceptible to either marginal breakdown or possible failure, including brittleness. Porosity, which usually exists to some extent in most amalgams, accelerates the corrosion of the material [14].
High-copper amalgam has a composition of 40–60 wt. % silver, 12–30 wt. % tin, 8.5–33 wt. % copper, 0.5–9 wt. % indium, and 1–3.5 wt. % zinc, and in some amalgam alloys, it has a trace amount of palladium (0.7 wt. %) [15]. Researchers have suggested that high-copper amalgam must have a minimum of 12 wt. % copper for the elimination of the γ2 phase, which causes an emission of a high concentration of mercury vapor from the alloy [16,17,18,19]. The development of high-copper amalgam by adding silver–copper eutectic particles to traditional silver–tin lathe-cut particles while dispersion hardening the alloy is reported to produce improved physical properties [20]. However, these are not the result of the dispersion hardening, as the silver–copper eutectic particles were big and too spaced out to inhibit dislocation movement. Instead, these improved properties were due to the formation of Cu6Sn5 η-phase [20,21,22]. The great affinity of tin for copper ensures that the γ2-phase is significantly reduced or partially eliminated in addition to generating visible improvements in physical properties such as increased strength, less tarnishing, corrosion resistance, and creep reduction.
Increased levels of weight percent copper present in dental amalgam alloy [23] enable the interdiffusion of copper into the elemental mercury upon trituration, thus initiating a reaction of Sn8Hg (γ2-phase). This γ2-phase is eliminated through the addition of copper to the alloy. Such methods of transformation occurred through changes in the Cu3Sn (ε-phase) [24,25]. The concept of enthalpy and the Gibbs free energy of the reaction and phase product are considered and utilized for the calculation of the energies released at each time interval during mercury vapor measurements. The two main phases that precipitate subsequent to trituration are Ag2Hg31-phase) and Sn8Hg (γ2-phase) [24]. Such precipitates initially coexist with the liquid mercury for a short period of time, approximately 10–15 min, and the mixture maintains a plastic consistency, which allows for the placement and shaping of the amalgam during its condensation into the tooth-prepared cavity. The metallic powder is made to be mixed in correct proportion with liquid mercury (consisting of 40–50 wt. %). During trituration, the mercury becomes supersaturated with the metallic atoms, thus leading to nucleation and the growth of the distinct phases which eventually precipitate from the alloy solution [26,27,28].
The alloy particles are manufactured in micro-cut, fine-cut, and coarse-cut particle sizes. The alloy generally develops phases in the form of binary phases, Ag-Sn (γ-phase), Ag-Sn-Cu (ternary phase), and Ag-tin-Cu-Zn (quaternary phase) [29]. The solubility of copper in the Ag3Sn (γ-phase) is only 1 wt. %. Therefore, excess copper forms the copper-rich phase. The amount and type of phase may vary due to the thermal processing, including the Cu3Sn (ε-phase) and the Cu6Sn5 (ζ-phase). High-copper admixed alloy contains a eutectic microstructure of an Ag-rich γ-phase and a copper-rich γ-phase [30]. Zinc is present in the low-copper amalgam powder, and its concentration exceeds the solubility limits of 1.6 wt. % Ag-Sn (β-phase), and 5.9% subsequently forms Cu5Zn8 [31,32].
During the amalgam aging process, changes in the composition and microstructural phases occur. The reaction of mercury and mercury-rich amalgam continues during the setting. High-copper amalgam produces a lesser amount of γ2-phase and transforms into the Cu6Sn5 phase. Amalgams that have been in place for many years usually present some transformation of the Ag2Hg31-phase), subsequently resulting in the Ag9Hg111-phase) [33]. This loss can occur due to three possible mechanisms:
  • Dissolution of mercury at the amalgam surface.
  • Evaporation from the exposed surface of the amalgam.
  • Migration of mercury to the interface of the remaining portion of the unreacted high-copper particles [34].
Two other phase changes generally seem to manifest and are likely to occur during the aging process besides those associated with corrosion, one of which is the reaction of the deleterious Sn8Hg (γ2-phase) phase with unreacted Cu3Sn to form the Cu6Sn eta-phase [24]. Such will occur in the copper alloy if there is any Sn8Hg (γ2-phase). The other likely change is that of the Ag2Hg31-phase) into Ag9Hg111-phase) [24]. This phenomenon has been observed in high-copper alloy surfaces where corrosion occurred. Diffusion of mercury frequently occurred during transformation of the alloy at room temperature during setting process. Diffusion of mercury vapor originates from the Ag2Hg3 γ1-rich mercury phase, which regularly transforms into the Ag9Hg11 β1 phase [34,35,36,37].
This study examines the microstructural changes of the material with advanced aging. Previous research has shown that when the material slowly undergoes oxidation, the mercury vapor levels decrease as a function of time and the oxidative process [37,38]. Such factors control the material properties, such as compressive strength, ductility, hardness, corrosion resistance, creep, and mercury vapor emission [39]. The research objective was primarily to determine and analyze the amount of mercury vapor released from each brand of dental amalgam, having different copper contents, with equal time duration of vapor measurement.

2. Materials and Equipment

Amalgam brands of capsules in spells of 400 mg, 600 mg, and 800 mg were obtained from commercial distribution dental supply companies such as Henry-Schein Dental (Melville, NY, USA; www.henryschein.com), Darby Dental Supply (Jericho, NY, USA; www.darbydental.com), and Patterson Dental Supply. The amalgam capsules obtained from each of the three suppliers are as follows: (1) Dispersalloy (11.8% Cu, regular-set, 2-spills, 600 mg, blue/white color capsule), (2) Permite C/SDI (15.4% Cu, regular set, 2-spills, 600 mg, purple/gray color capsule), (3) Valliant PHD (20% Cu, regular set, 2-spills, 600 mg, blue color capsule), (4) Megalloy EZ (25% Cu, 56% silver, regular set, 2-spills, 600 mg, purple color capsule), (5) Tytin FC/Kerr (26% Cu, regular set, 2-spills, 600 mg, blue/white color capsule), (6) Tytin/Kerr (28% Cu, regular set, 2-spills, 600 mg, white/yellow color capsule), Tytin/Kerr (28% Cu), (7) Contour/Kerr (31% Cu, regular-set, 2-spills, 600 mg, brown/gray color capsule), (8) Sybralloy/Kerr (33% Cu. regular-set, 2-spills, 600 mg, green/gray colored capsule).
In order to measure the concentration of mercury vapor from freshly condensed dental amalgams of various brands and the energy released at each stage of measurement, a Jerome J505 Mercury Vapor Analyzer (Arizona Instrument, Chandler, AZ, USA) was used during the experimental procedure. The Jerome J505 Instrument works on the principle of drawing an inflow of saturated air with particulates via a built-in pump. The air is made to flow over a gold-metallic strip at a warm temperature. Since gold has a high affinity for mercury atoms, the mercury atoms in the saturated air enter the instrument through an orifice by way of a 12-inch plastic tubing attached to the instrument [40].
The standard unit range for the Jerome J505 Mercury Vapor Analyzer is ng/m3 (50 to 500,000), μg/m3 (0.05 to 500), and mg/m3 (0.00005 to 0.50000). The instrument automatically computes statistical parameters such as the variance, standard deviations, and percentage error in the data obtained. Percentage error indicated ±0.05%, each successive measurement demonstrated similarities in results, and each measurement was obtained under normal conditions [40]. In order to understand the phase purity and orientation of the Dispersalloy and Sybralloy, X-ray diffraction analysis was performed using the Bruker D8 (Billerica, MA, USA).

3. Methodology

The research investigated various samples of dental amalgam prepared first by trituration followed by condensation into samples with diameters of 10 mm andthicnkesses of 4 mm. For this research, eight of the most common brands of dental amalgams (having varying powder compositions) were obtained from commercial manufacturers: Dispersalloy—11.8% Cu, Permite C/SDI—15.4% Cu, Valliant PHD—20% Cu, Megalloy EZ—25% Cu, Tytin FC/Kerr—26% Cu, Tytin/Kerr—28% Cu, Contour/Kerr—31% Cu, Sybralloy/Kerr—33% Cu.
Each sample was polished using rough, medium, and smooth-grade Emory polishing paper and was measured under similar environmental conditions and operating speed (temperature between 20–30 °C at 1 atm). These amalgams were distributed in capsules, and inside this capsule was a thin plastic/polymer membrane that separates the metallic powder on one side from the liquid mercury on the other side. Prior to trituration, the amalgam capsule was inserted between the two vibrating prongs of the Zenith dental variable speed amalgamator, which features a torque motor, having the choice of three speed settings: high (4800 rpm), medium (4200 rpm), and low (3600 rpm). The amalgamator was connected to a 120-vold power supply. At the onset of electric power, the amalgam capsule was allowed to vibrate at this high speed for 15 s for the trituration process of mixing.
Prior to initiating measurement, the Mercury Vapor Analyzer was warmed-up for about 10 min, and a test sample reading was obtained in order to verify the accuracy of the measurements. Amalgam capsules were then condensed and prepared in uniform size. The samples were inserted into a Stony Lab 250 mL borosilicate glass (Nesconset, NY, USA). A 12-inch length of plastic tubing with a diameter of 4 mm was attached to the Stony Lab glass, while the other end was attached to the J505 instrument. At the onset of operations, saturated ambient air containing mercury atoms was drawn through a 12-inch plastic tube with a diameter of 4 mm attached to the 250 mL flask. Saturated air samples containing mercury atoms were drawn into the instrument by means of a built-in electrical pump located inside the instrument. The normal flow rate of saturated air is 1 L/min. The sample air then flows through a scrubber filter and directly into the sample cell located inside the instrument; the entrance of the sample cell is controlled by a one-way valve to prevent back-flow.
Six weeks subsequent to sample preparation, the hardness of each sample was measured using the Microvickers Hardness Tester, Model M-400-H1 (Mitutoyo, Kawasaki, Japan).
Thermodynamic calculation methods were applied in determining the energies released from each sample’s stage of concentration measurement (see Equation (1) [29]. All of the measurements were obtained at room temperature during the research experiment (temperature between 20–30 °C at 1 atm).
The study also examined the microstructural changes of the material with advanced aging by analyzing the X-ray diffraction pattern obtained from both the Dispersalloy and the Sybralloy using the D8 Bruker.

4. Results and Discussions

Table 1 below shows the results recorded from the Mercury vapor analyzer. In accordance with the results, and as demonstrated by Figure 1 below, the amalgam samples released a significant amount of mercury vapor. The Dispersalloy (11.8% Cu) released the highest concentration of mercury vapor, while Sybralloy (33% Cu) released the lowest amount of vapor. The vapor release measurement was conducted starting at zero seconds followed by an interval of 20 s up to 2800 s.

4.1. Mercury Vapour Concentration

The results from Table 1 show the decrease in mercury vapor released with the passage of time and with the increase in weight percent copper in each sample.
At the starting time (zero seconds) of the mercury vapor measurement of each alloy, the mercury vapor level for the Dispersalloy (11.8% Cu) gave a value of 846 kg/m3 compared to the concentration value of 796 kg/m3 for the Permite C/SDI (15.4% Cu) brand. For the same starting time for the Permite C/SDI (15.4% copper), the mercury released was much lower in value than that of the Dispersalloy amalgam. This is a verification that the amalgam alloy with the lowest percentage of copper will release the greatest concentration of mercury vapor, because the increased amount of copper added to the alloy will suppress the release of mercury vapor. Such evidence of the mercury vapor released from each amalgam alloy can be seen in the plots of Figure 1 and Figure 2, showing the mercury released for each alloy at intervals of 50 s, 200 s, and 350 s.
As observed from the plot of Figure 2, both the mercury vapor levels released from the Contour (31% Cu) amalgam and the Sybralloy (33%) amalgam appear to be about the same. Both alloys have approximately the same amount of copper content in their compositions, thus, each alloy releases that amount of mercury vapor in very close proximity. For the high-copper amalgam alloys such as Contour and Sybralloy, the much-increased copper content serves to eliminate the γ2 phase which is responsible for inducing corrosion and tarnishing within the alloy [41,42].

4.2. X-ray Diffraction Analysis

The X-ray diffraction pattern of Dispersalloy (11.8% Cu) and Sybralloy (33% Cu) shows the relative comparison of the Sn8Hg (γ2-phase), of which a greater amount is present in the Dispersalloy plot, as is demonstrated in Figure 1. The major phase in each the Dispersalloy (11.8% Cu) and Sybralloy (33% Cu) brands is the silver–tin (γ-phase), the strongest phase having the tallest peak. The mercury vapor is released during the γ1-phase (Ag2Hg3) and mostly from Sn8Hg (γ2-phase), which is predominant in the low-copper alloy. The Sybralloy (33% Cu) from Figure 1 showed a reduction in the γ2-phase due to the higher weight percent copper in that alloy. Interestingly, 15% vol. of the matrix phase is composed of Sn8Hg (γ2 phase), which is usually a stable phase [32]. This means that the lesser amount of copper enables strong bonding and is brooched in the Cu3Sn ε-phase or Cu6Sn5 η-prime phase [32,33].
The X-ray diffraction image for Dispersalloy as shown in Figure 3 shows a large accumulation of the γ2-phase, and the X-ray diffraction for Sybralloy showed a lesser amount of the γ2- phase, thus indicating a lesser amount of mercury vapor released from Sybralloy than from Dispersalloy and for the other amalgam samples. Previous research has shown that the material slowly undergoes oxidation, and the mercury vapor levels decrease as a function of time and the oxidative process [37]. Such factors control the material properties, such as compressive strength, ductility, hardness, corrosion resistance, creep, and mercury vapor emission [39].
Attempts have been made to reduce the γ2 phase by increasing the copper content in the alloy, effectively above 13% [1]. It is established that the early full strength of the amalgam is achieved within one hour of placement in the prepared tooth cavity [5]. The setting reaction of this alloy is the same as the reaction for the conventional alloy [10]. After the formation of the γ2 phase, there is a reaction between this and the silver–copper component, leading to the formation of the copper–tin phase and γ1 phase [10,18].
The results of the Vickers hardness measurement as shown in Table 2 above for each of the amalgam samples, showed that the hardness of the amalgam increased with the increase in copper content within each of the samples. Such phenomena are, however, independent of time.

4.3. Analysis of Energy Released

The energy of formation given off or generated due to the release of mercury vapor can be determined by the following thermodynamic method (Equation (1)):
E = C   ×   H M
where:
  • E = Energy of reaction due to vapor released.
  • C = Recorded concentration of mercury vapor released.
  • H = Energy given off when one atom of mercury is released (32 kilojoules per mole).
  • M = Molecular weight of mercury (200.59 g per mole)
The energies given off from each of the amalgams at each stage of measurement are determined from the calculations as indicated in Equation (1). A decrease in energy is observed with time; therefore, it can be concluded that the energy is proportional to the concentration of mercury vapor released from the amalgam, regardless of the brand. Also, the higher the copper content present in the alloy, the less mercury vapor is released from the amalgam.
In accordance with the results generated from the plots, it is determined that the amalgam brands with higher copper percentages tend to release a lower concentration of mercury vapor within a shorter time interval, accounting for the steeper gradient of higher copper levels within the alloy. The energy of formation is indeed lower for the alloy having the higher copper percentage in the alloy composition.
The results of the experimental data, as listed in Table 1, give the values of the concentrations of mercury vapor as given off by various amalgam alloys. From the Dispersalloy (11.8% copper) brand of dental amalgam to the Sybralloy (33% copper), the measurement of mercury vapor shows a decrease, indicating that vapor release is directly proportional to the copper content of the alloy. Additionally, the mercury vapor decreases with an increase in time and attains a somewhat steady-state value. Such steady-state values will be further achieved due to the oxidation of the alloy with the passage of time.
Both the Contour (31% Cu) and Sybralloy (33% Cu) brands of amalgam demonstrate the lowest energy of formation, which supports the conclusion that the highest copper content in the alloy produces the least concentration of mercury vapor released from the amalgams. This research has proven the desired results of the experiment. As shown, the hardness of the amalgams increases with increasing copper percentage. The corresponding plot is shown in Figure 4, which shows a graphical representation of the phenomena.
Further investigation regarding the amalgam alloy is required for designing alloys with optimum compositions. In recent years, the American public has become more concerned and ambivalent about the continued mercury vapor release from dental amalgam. Presumably, as more knowledge of this material propagates, there is a high chance that amalgam could become obsolete due to the concerns of the mercury released from such material.
The energy released versus the time plot was tabulated as shown in Figure 5. The energies given off from each of the amalgams were determined from calculations as shown in Equation (1).

5. Conclusions

  • A systematic study has been performed to determine the release of mercury vapor from the eight most common brands of dental amalgam. The release of Hg vapor in order of decreasing amount is found to be as follows: Disperesalloy Brand, Permite C/SDI, Valliant PHD, Megalloy EZ, Tytin FC, Tytin/kerr, Contour, and Sybralloy.
  • The hardness of the amalgam is inversely proportional to the mercury vapor released from the alloy (i.e., hardness increases with decreasing copper percentage).
  • X-ray diffraction confirms larger accumulations of γ2-phase subsequent to trituration.
  • The amount of energy required for the removal of mercury atoms during vapor release decreases with increasing time duration.
  • A new dental amalgam alloy can be achieved by possibly introducing a new metal to the existing alloy, such as titanium powder.
In brief, this study concludes that the copper content in amalgam is correlated with the amount of Hg vapor released from the alloy. The low-copper amalgam showed higher releases of Hg vapor. In order to manufacture an improved amalgam alloy, either a reduced quantity of mercury or an increased quantity of copper should be considered when designing the alloy composition, along with the introduction of new material if possible. More research and investigations need to be made into the modification of this alloy without significantly affecting its properties.

Author Contributions

Conceptualization, R.M. and D.K.; Methodology, R.M. and Z.X.; Validation, R.M., Z.X., I.C.-O. and S.C.; Investigation, R.M.; Data curation, R.M.; Writing—review & editing, I.C.-O., S.C. and D.K.; Project administration, D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NSF-Partnership for Research and Education in Materials (PREM) grant number DMR-2122067.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

R.M. acknowledges the financial support he received from the NSF-Partnership for Research and Education in Materials (PREM) grant number DMR-2122067 at North Carolina A and T State University.

Conflicts of Interest

The authors have no conflict of interest.

References

  1. Okabe, T.; Mitchell, R.J. Setting reactions in dental amalgam, part 2. The kinetics of amalgamation. Crit. Rev. Oral Biol. Med. 1996, 7, 23–35. [Google Scholar] [CrossRef] [PubMed]
  2. Aaseth, J.; Hilt, B.; Bjørklund, G. Mercury exposure, and health impacts in dental personnel. Environ. Res. 2018, 164, 65–69. [Google Scholar] [CrossRef] [PubMed]
  3. De Oliveira, M.T.; Pereira, J.R.; Ghizoni, J.S.; Bittencourt, S.T.; Molina, G.O. Effects from exposure to dental amalgam on systemic mercury levels in patients and dental school students. Photomed. Laser Surg. 2010, 28, S111–S114. [Google Scholar] [CrossRef] [PubMed]
  4. UNEP. Global Mercury Supply, Trade, and Demand; United Nations Environment Programme, Chemicals and Health Branch: Geneva, Switzerland, 2017; p. 96. [Google Scholar]
  5. Mante, F.K.; Chern-Lin, J.H.; Mante, M.O.; Greener, E.H. The effect of noble metals on the mechanical properties of dispersed phase dental amalgam. J. Oral Rehabil. 1998, 25, 279–284. [Google Scholar] [CrossRef]
  6. Craig, R.G.; Powers, J.M.; Wataha, J.C. Dental Materials, Properties and Manipulations, 7th ed.; Mosby-Year Brook Inc.: St. Louis, MO, USA, 2000. [Google Scholar]
  7. Echeverria, D.; Woods, J.S.; Heyer, N.J.; Rohlman, D.; Farin, F.M.; Li, T.; Garabedian, C.E. The association between a genetic polymorphism of coproporphyrinogen oxidase, dental mercury exposure and neurobehavioral response in humans. Neurotoxicol. Teratol. 2006, 28, 39–48. [Google Scholar] [CrossRef]
  8. Khwaja, M.A.; Abbasi, M.S. Mercury poisoning dentistry: High-level indoor air mercury contamination at selected dental sites. Rev. Environ. Health 2015, 29, 29–31. [Google Scholar] [CrossRef]
  9. Bernhoft, R.A. Mercury toxicity and treatment: A review of the literature. J. Environ. Public Health 2012, 2012, 460508. [Google Scholar] [CrossRef]
  10. Okabe, T.; Mitchell, R.J. Setting Reactions in Dental Amalgam Paper 2. The Kinetics of Amalgam. Crit. Rev. Oral Biol. Med. 1992, 7, 23–35. [Google Scholar]
  11. Ferracane, J.L.; Hanawa, T.; Okabe, T. Effectiveness of oxide films in reducing mercury release from amalgams. J. Dent. Res. 1992, 71, 1151–1155. [Google Scholar] [CrossRef]
  12. Gayler, M.L. The constitution of the alloys of silver, tin, and mercury. J. Inst. Metals 1937, 60, 379. [Google Scholar]
  13. Jamil, N.; Baqar, M.; Ilyas, S.; Qadir, A.; Arslan, M.; Salman, M.; Ahsan, N.; Zahid, H. Use of Mercury in dental silver amalgam: An occupational and environmental assessment. BioMed Res. Int. 2016, 2016, 6126385. [Google Scholar] [CrossRef]
  14. Dhar, V.; Hsu, K.L.; Coll, J.A.; Ginsberg, E.; Ball, B.M.; Chhibber, S.; Johnson, M.; Kim, M.; Modaresi, N.; Tinanoff, N. Evidence-base update of pediatric dental restoration procedures; dental materials. J. Clin. Pediatr. Dent. 2015, 39, 303–310. [Google Scholar] [CrossRef]
  15. Kopperud, S.E.; Tveit, A.B.; Gaarden, T.; Sandvik, L.; Espelid, I. Longevity of posterior dental restoration and restorations and reasons for failure. Eur. J. Oral Sci. 2012, 120, 539–548. [Google Scholar] [CrossRef]
  16. Kall, J.; Just, A.; Aschner, M. What is the risk? Dental amalgam, mercury exposure, and human health risks throughout the life span. In Epigenetics, the Environment, and Children’s Health Across Lifespans; Hollar, D., Ed.; Springer: Cham, Switzerland, 2016; pp. 159–206. [Google Scholar]
  17. Koral, S.M. Mercury from dental amalgam: Exposure and risk assessment. Compend. Contin. Educ. Dent. 1995, 34, 138–140. [Google Scholar]
  18. Davies, R.A.; Ardalan, S.; Mu, W.H.; Tian, K.; Farsaikiya, F.; Darvell, B.W.; Chass, G.A. Geometric, electronic, and elastic properties of dental silver amalgam γ-(Ag3Sn), γ1-(Ag2Hg3), γ2-(Sn8Hg) phases, comparison of experiment and theory. Intermetallics 2010, 18, 756–760. [Google Scholar] [CrossRef]
  19. Ghatee, M.H.; Karimi, H.; Shekoohi, K. Structural, Mechanical, and thermodynamical properties of silver amalgam filler: A Monte Carlo simulation study. J. Mol. Liq. 2015, 211, 96–104. [Google Scholar] [CrossRef]
  20. Szurgot, K. Microstructure of Ag-Cu-Pd dispersed phase amalgam. J. Dent. Res. 1981, 60, 481. [Google Scholar]
  21. Okabe, T.; Mitchell, R.; Butts, M.B.; Wright, A.H.; Fairhurst, C.W. A study of high copper amalgams. I. A Comparison of amalgamation on a hg-plated high copper alloy tablets. J. Dent. Res. 1978, 57, 759–767. [Google Scholar] [CrossRef]
  22. Eames, W.B. Preparation, and condensation of amalgam with low mercury alloy ratios. J. Am. Dent. Assoc. 1959, 58, 78–83. [Google Scholar] [CrossRef]
  23. Demaree, N.C.; Taylor, D.F. Properties of dental amalgam from spherical alloy particles. J. Dent. Res. 1962, 41, 890–906. [Google Scholar] [CrossRef]
  24. Innes, D.B.K. Dispersion strengthened amalgam. Can. Dent. Assoc. J. 1963, 29, 587–593. [Google Scholar]
  25. Mayhew, R.B.; Schmeltzer, L.D.; Pierson, W.P. Effects of polishing on the marginal integrity of high copper amalgam. Oper. Dent. 1968, 11, 8–13. [Google Scholar]
  26. Okabe, T.; Ohmoto, K.; Nakajima, H.; Woldu, M.; Ferracane, J.L. Effect of Pd and in on mercury evaporation. Dent. Mater. J. 1987, 16, 191–199. [Google Scholar] [CrossRef] [PubMed]
  27. Svare, C.W. Dental Amalgam Related Mercury Vapor Exposure. CDA J. 1984, 12, 35–56. [Google Scholar]
  28. Vimy, M.J.; Lorscheider, F.L. Intra-oral Air Mercury Released from Dental Amalgams. J. Dent. Res. 1985, 64, 1069–1071. [Google Scholar] [CrossRef]
  29. Masi, J.V. Corrosion of amalgam in restorative materials: The problem and the promise. In Status Quo and Perspective of Amalgam and Other Dental Materials; Friberg, L., Schrauzer, G.N., Eds.; Thieme Verlag: Stuttgart, Germany, 1995; in press. [Google Scholar]
  30. Okabe, T.; Mitchell, R.J.; Butts, M.B.; Ferracane, J.L. Microstructural changes on the surface of dental amalgam during aging. In Microstructural Science; International Metallographic Society: Detroit, MI, USA, 1986; Volume 13, pp. 177–195. [Google Scholar]
  31. Ferne, C.A.; Asgar, K.; Peyton, F.A. Microstructure of Dental Amalgam. J. Dent. Res. 1965, 44, 1002–1012. [Google Scholar] [CrossRef]
  32. Colon, P.; Pradelle-Plasse, N.; Galland, J. Evaluation of the long-term corrosion behavior dental amalgams: Influence of palladium addition and particle morphology. J. Dent. Mater. 2003, 19, 232–239. [Google Scholar] [CrossRef]
  33. Koike, M.; Ferracane, J.L.; Adey, J.D.; Fujii, H.; Okabe, T. Initial mercury evaporation from experimental Ag-Sn-Cu amalgams containing Pd. Biomaterial 2004, 25, 3147–3153. [Google Scholar] [CrossRef]
  34. Lin, J.C.; Marshall, S.J.; Marshall, G.W. Microstructure of high copper amalgams after corrosion in various solutions. J. Dent. Mater. 1987, 3, 176–181. [Google Scholar]
  35. Sun, Y.; Yu, H.; Kesim, M.T.; Alpay, S.P.; Aindow, M. Microstructural Stability, Defect Structures and Deformation Mechanisms in a Ag3Sn/Cu3Sn Alloy. J. Mater. Sci. 2017, 52, 2944–2956. [Google Scholar] [CrossRef]
  36. Powell, L.V.; Johnson, G.H.; Bales, D.J. Effect of Admixed Indium on Mercury Vapor Release from Dental Amalgam. J. Dent. Res. 2016, 68, 1231–1233. [Google Scholar] [CrossRef]
  37. Pihl, C.F.; Beasley, W.M. Compounds Formed in Silver Dental Amalgam. J. Dent. Res. 2016, 47, 418–426. [Google Scholar] [CrossRef]
  38. Donly, K.J.; Sasa, I.S. Dental Materials Pediatric Dentistry, 6th ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 293–303. [Google Scholar] [CrossRef]
  39. Aerts, A.; Danaci, S.; Prieto, B.G.; Van den Bosch, J.; Neuhausen, J. Evaporation of mercury impurity from liquid lead–bismuth eutectic. J. Nucl. Mater. 2014, 448, 276–281. [Google Scholar] [CrossRef]
  40. Arizona Instrument Inc. Jerome J505 Mercury Vapor Analyzer; Operation Manual; Arizona Instrument LLC: Chandler, AZ, USA, 2015; pp. 1–28. Available online: https://www.raecorents.com/amfile/file/download/file/238/product/633/ (accessed on 2 March 2023).
  41. Mitchell, R.J.; Okabe, T. Setting reactions in dental amalgam, Part 1. Phases and microstructures between one hour and one week. Crit. Rev. Oral Biol. Med. 1996, 7, 12–22. [Google Scholar] [CrossRef]
  42. Halbach, S.; Welzl, G.; Kremers, L.; Willruth, H.; Mehl, A.; Wack, F.X.; Hickel, R.; Greim, H. Steady-state transfer and depletion kinetics of mercury from amalgam fillings. Sci. Total Environ. 2000, 259, 13–21. [Google Scholar] [CrossRef]
Figure 1. Decrease in the mercury vapor concentrations of dental amalgam samples as a function of time for the first 1000 s.
Figure 1. Decrease in the mercury vapor concentrations of dental amalgam samples as a function of time for the first 1000 s.
Materials 16 03452 g001
Figure 2. Comparison of the amount of mercury vapor released from eight amalgam samples with different Cu contents (indicated in the figures) as a function of time: (a) Disperse and Permite, (b) Valiant PHD and Megalloy EZ, (c) Tytin FC/Kerr and Tytin/Kerd, and (d) Controur and Sybralloy dental amalgam.
Figure 2. Comparison of the amount of mercury vapor released from eight amalgam samples with different Cu contents (indicated in the figures) as a function of time: (a) Disperse and Permite, (b) Valiant PHD and Megalloy EZ, (c) Tytin FC/Kerr and Tytin/Kerd, and (d) Controur and Sybralloy dental amalgam.
Materials 16 03452 g002
Figure 3. XRD pattern for the Dispersalloy (11.8% Cu) and Sybralloy (33% Cu) dental amalgams.
Figure 3. XRD pattern for the Dispersalloy (11.8% Cu) and Sybralloy (33% Cu) dental amalgams.
Materials 16 03452 g003
Figure 4. Increase in the Vickers hardness number as a function of copper percentage in the alloy.
Figure 4. Increase in the Vickers hardness number as a function of copper percentage in the alloy.
Materials 16 03452 g004
Figure 5. Energy of formation of alloys containing different amounts of Cu as a function of time.
Figure 5. Energy of formation of alloys containing different amounts of Cu as a function of time.
Materials 16 03452 g005
Table 1. Experimental data for mercury vapor concentration (kg/m3) released from eight brands of dental amalgam samples.
Table 1. Experimental data for mercury vapor concentration (kg/m3) released from eight brands of dental amalgam samples.
Time (s)Disperse Alloy (kg/m3)Permite C/SDI (kg/m3)Valiant PHD (kg/m3)Megalloy EZ (kg/m3)Tytin FC/Kerr (kg/m3)Tytin/Kerr (kg/m3) Contour (kg/m3)Sybralloy (kg/m3)
11.8% Cu15.4% Cu20% Cu25% Cu26% Cu28% Cu31% Cu(33% Cu)
0846.085796.8452788.7453654.4776651.6575615.6716513.0673509.6588
20838.5543790.5823784.0675651.0781647.4575612.7408510.4512507.6844
40821.8564788.5934781.8564648.4752642.8862608.8856507.8546504.8556
60796.6544785.6743779.5686645.3237640.5644602.5639505.4364501.5342
80787.0785778.5879774.9676643.9465638.5997598.4522499.5089497.8673
100780.6533765.8547764.6354641.7498635.7675595.4965496.2395495.9536
120773.851753.6457750.6793639.4652632.7673591.2894490.2007491.8945
140767.3113749.4512744.8476636.3649629.5088588.2098487.9915489.6577
160756.4582745.463742.9806632.8451625.6554585.7898485.9067485.9676
180748.9329738.8486734.5864631.9463622.8789583.8456484.773483.7462
200734.831730.6582729.5655629.4856619.7769580.5487482.9566481.6766
220723.0086720.7678719.8867626.8354617.3744578.3008480.6876478.5653
240706.5386703.5541700.8723624.0376614.8842575.3065478.1153475.7674
260684.1459679.7895675.9665621.2743612.3428573.0678473.5972473.8665
280675.8653673.8823671.5434619.8211610.8452569.6458471.3412471.8604
300658.6462654.6008650.7685617.3845608.4768567.7342465.8456469.6997
320623.0678621.0804618.4437616.3329606.5646563.9647463.698464.0767
340595.9563591.4318589.4878585.5875583.5453560.7456461.9438461.5682
360590.5342589.6074586.9652583.7677581.4436554.7439457.8348458.0634
380586.6308583.3116580.5896577.3458575.3876551.8856453.1195454.7683
400577.8899570.5078565.5768561.6455559.6763547.4865451.7344451.7554
Table 2. Hardness for each amalgam brand with corresponding percent copper in the alloy.
Table 2. Hardness for each amalgam brand with corresponding percent copper in the alloy.
AlloyPercentage CopperHardness
Disperse Alloy11.864HV200
Permite 15.480HV200
Valliant PHD2093HV200
Megalloy EZ25100HV200
Tytin FC26100.8HV200
Tytin/Kerr28101.7HV200
Contour31112HV200
Sybralloy33129HV200
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

Moxon, R.; Xu, Z.; Chris-Okoro, I.; Cherono, S.; Kumar, D. Determination and Calculations of Mercury Vapor Concentration and Energy Released from Freshly Condensed Dental Amalgams Having Various Copper Percentages within the Alloy. Materials 2023, 16, 3452. https://doi.org/10.3390/ma16093452

AMA Style

Moxon R, Xu Z, Chris-Okoro I, Cherono S, Kumar D. Determination and Calculations of Mercury Vapor Concentration and Energy Released from Freshly Condensed Dental Amalgams Having Various Copper Percentages within the Alloy. Materials. 2023; 16(9):3452. https://doi.org/10.3390/ma16093452

Chicago/Turabian Style

Moxon, Ryan, Zhigang Xu, Ikenna Chris-Okoro, Sheilah Cherono, and Dhananjay Kumar. 2023. "Determination and Calculations of Mercury Vapor Concentration and Energy Released from Freshly Condensed Dental Amalgams Having Various Copper Percentages within the Alloy" Materials 16, no. 9: 3452. https://doi.org/10.3390/ma16093452

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